Sie sind auf Seite 1von 17

BIOREACTORS, CONTINUOUS STIRRED-TANK REACTORS 353

172. G. Rossi, Biohidrometallurgy, McGraw-Hill, Hamburg, Ger- BIOREACTORS, CONTINUOUS STIRRED-TANK


many, 1990. REACTORS
173. B.C. Smith and D.R. Skidmore, Biotechnol. Bioeng. 35, 483–
491 (1990). MICHAEL E. LASKA
174. A. Assa and R. Bar, Biotechnol. Bioeng. 38, 1325–1330 CHARLES L. COONEY
(1991). Massachusetts Institute of Technology
175. J. Schmidt, R. Nassar, and A. Lubbert, Chem. Eng. Sci. 47, Cambridge, Massachusetts
2295–2300 (1992).
176. J.B. Snape and N.H. Thomas, Biotechnol. Bioeng. 40, 337–
345 (1992). KEY WORDS
177. N.M.G. Oosterhuis, Ph.D. Thesis, Dept. of Chemical Tech-
nology, Univ. of Technology, Delft, The Netherlands, 1984. Chemostat
178. N.W.F. Kossen and N.M.G. Oosterhuis, in N.J. Rhem and J. Continuous culture
Reed eds. Biotechnology VCH, Weinheim, Germany, 1985. Continuous stirred-tank reactor (CSTR)
179. A.P.J. Sweere, K.C.A.M. Luyben, and N.W.F. Kossen, En-
zyme Microb. Technol. 9, 386–392 (1987).
180. C. Sola and F. Godia, in J.A. Asenjo, J.C. Merchuk eds., Bio- OUTLINE
reactor System Design, Dekker, New York, 1994.
181. F.H. Johnson, H. Eyring, and M.J. Polissar, The Kinetic Ba- Introduction
sis of Molecular Biology, Wiley, New York, 1954. Definitions
182. D.G. Jordan, Chemical Process Development, Volume 6, Wi- Strategy for CSTR Analysis
ley, New York, 1968.
Reaction Kinetics
183. H. Taguchi, T. Imanaka, S. Teramoto, M. Takatsu, and M.
Sato, J. Ferment. Technol. 46, 823–830 (1986). Cell Growth
184. H. Takei, K. Misusawa, and F. Yoshida, Ferment. Technol. Enzymes
53, 151–156 (1975). Mass Transfer
185. J.A. Roels, Energetics and Kinetics in Biotechnology, Elsev- Introduction
ier, Amsterdam, 1983.
Variations on the Single CSTR
186. B. McNeil and B. Kristiansen, Biotechnol. Lett. 9, 101–107
(1987). Single CSTR with Recycle
187. P.M. Doran, Adv. Biochem. Eng. Biotechnol. 48, 115–168 CSTRs in Series
(1993). Nomenclature
188. D.J. Pollard, A.P. Ison, P. Ayazi Shamlou, and M.D. Lilly, The Bibliography
Examination of Bioreactor Heterogeneity with Rheological
Different Fermentation Broths, Kluwer, Dordrecht, The
Netherlands, 1994, pp. 163–170. INTRODUCTION
189. F. Potucek, Collect. Czech. Chem. Commun. 55, 981–986
(1990). Definitions
190. P.D. Gaspillo and S. Goto, J. Chem. Eng. Jpn. 24, 680–682 A continuous stirred-tank reactor (CSTR) is defined as an
(1991). agitated vessel with continuous addition and removal of
191. Y. Chisti, M. Kasper, and M. Moo-Young, Can. J. Chem. Eng. material and energy. The CSTR is one of the basic contin-
68, 45 (1990). uous reactor types widely used in the chemical process in-
192. S.Q. Zhou, L.M. Tang, and K. Schugerl, J. Biotechnol. 28, dustries because of its amenability to process control and
165–177 (1993). scale-up, although in biotechnology applications the CSTR
193. Y. Bando, M. Nishimura, H. Sota, S. Suzuki, and N. Kawase, is used more often as a research tool than as a production
Chem. Eng. Sci. 47, 3371–3378 (1992). technology. An idealized, well-mixed CSTR can be modeled
194. Y. Bando, M. Nishimura, A. Hayaashi, S. Hiura, S. Indo, and as having no spatial variations in temperature, concentra-
A. Idota, J. Chem. Eng. Jpn. 28, 225–227 (1995). tion, fluid properties, or reaction rates. Thus, the proper-
195. T.K. Ghosh, B.R. Maiti, and B.C. Bhattacharyya, Biotechnol. ties of the exit stream may be considered the same as those
Tech. 7, 301–307 (1993). throughout the vessel. Although such ideal mixing is never
196. C. Schlotelburg, M. Gluz, M. Popovic, and J.C. Merchuk, Int. observed, the vessel is designed to provide good mixing
Symp. on Bubble Columns, Kyoto, Nov. 1997. through selection of operating conditions and vessel, baffle,
and impeller geometries. The stirred tank used in contin-
uous bioprocesses is similar to that used in batch biopro-
See also BIOREACTORS, CONTINUOUS STIRRED-TANK cesses, with the exception that the CSTR likely has an
REACTORS; BIOREACTORS, FLUIDIZED-BED; overflow or other level control device. Oxygen can be intro-
BIOREMEDIATION; ENZYMES, IMMOBILIZATION duced into the vessel by sparging through inlets at the base
METHODS; FERMENTATION MONITORING, DESIGN AND of the vessel, where impellers then disperse the bubbles.
OPTIMIZATION; MASS TRANSFER; RHEOLOGY OF Vessel jacketing or internal cooling coils provide a means
FILAMENTOUS MICROORGANISMS, SUBMERGED for heat transfer. Continuous systems that are not agitated
CULTURE; SCALE-UP, STIRRED TANK REACTORS. vessels often are modeled as CSTRs when their behavior
354 BIOREACTORS, CONTINUOUS STIRRED-TANK REACTORS

approximates that of the ideal CSTR. CSTRs are also slowly, with typical doubling times between 18 and 48 h.
known as backmix reactors, continuous-flow stirred-tank Because of their larger size and lack of a protective cell
reactors (CFSTRs), or chemostats, when used for cell wall, mammalian cells are particularly sensitive to the
growth. fluid shear in the vessel as well as to the osmotic pressure
of the medium. Mammalian cells grow over a narrow range
Strategy for CSTR Analysis of osmotic/pressures and pH and typically have more com-
plex nutritional requirements than bacteria and yeasts do.
The performance and analysis of a CSTR is based on the Cell division in yeast occurs primarily by budding, with
material and energy conservation balances and the under- typical doubling times between 1 and 3 h. The budding
lying processes governing the reaction kinetics. Because of process leads to a mother and a daughter cell, each having
the large variety of CSTR applications and limited space different growth rates and cell surface characteristics. Un-
for discussion in this article, a brief outline of the steps in like bacterial cultures, yeast cell populations have a broad
systems analysis will be beneficial in understanding any and time-varying distribution of ages and properties that
CSTR-based process. may influence the formation of a desired product.
The strategy for analyzing CSTR performance first re- Mycelial growth occurs in molds, actinomycetes, and
quires defining the problem statement and goals. The sec- some yeasts by a process of hypha chain elongation and
ond step is system identification, which includes defining branching. Mycelial cultures also are characterized by a
the system boundaries and the interactions between the distribution of ages, with younger cells located at the hy-
system and its environment across the system boundaries. phal tips. The hyphae form intertwined cellular strands,
The system could be a cell, the fluid in the reactor, or an or mycelia, that increase the broth viscosity and lead to
entire bioprocessing plant. The third step is to identify the nonideal fluid mixing. High broth viscosity can be problem-
state variables that characterize the system. During the atic for process monitoring and control, cell separation and
course of the analysis, new state variables may be identi- recycle, and oxygen and heat transfer. Furthermore, fluid
fied and added to the original list. The fourth step is to shear in the vessel can cause hyphal breakage and the for-
characterize the state of the system using material and mation of denser, more highly branched pellets or flocs.
energy balances that account for the accumulation of mass Cells in these pellets may be exposed to different micro-
and energy. In a general balance for a particular quantity, environments because of mass transfer limitations, which
the rate of accumulation of that quantity in the system is can vary with culture conditions and influence important
equal to the net influx of the quantity across the system cell properties.
boundaries plus the rate at which the quantity is gener- Viral growth initially requires the infection of a host
ated. Separate material balances are written for the re- cell, which occurs by attachment of the virus to the cell
actants, products, and the catalyst (e.g., cells or enzymes). surface and injection of viral nucleic acids into the cell in-
The final step is to calculate performance metrics and re- terior. New viruses are constructed from biological mole-
visit the assumptions to determine the conditions under cules synthesized by the host cell under the direction of the
which they are valid. viral genome. Viral nucleic acid is replicated many times
(e.g., 500) and encapsulated in coat proteins to form a
REACTION KINETICS large number of new viral particles. Viral growth can pro-
ceed to one of two phases. In the lytic cycle, the host cell
This section discusses the theory governing ideal CSTR will lyse or break open and release infectious viral parti-
performance in two important bioprocess applications (cell cles, whereas in the lysogenic cycle, the viral DNA will be
growth and enzyme reactions), the underlying assump- integrated into the host cell DNA and the host cell will
tions of the theory, and some practical aspects associated continue to reproduce normally.
with CSTR operation. CSTRs used for cell growth are commonly referred to as
chemostats or turbidostats, depending on the strategy
used to control the vessel environment. The most common
Cell Growth
arrangement is the chemostat (1–3), in which the medium
Introduction. Cell growth occurs in response to the en- fed to the vessel is designed so that all but a single nutrient
vironment. It is useful to classify growth in four mecha- essential for growth are present in excess of the cells’ re-
nistic categories and identify those features relevant to the quirements. Any nutrient necessary for growth can be used
analysis of CSTRs. These classes are fission, budding, my- to control the size of the cell population in the vessel, mak-
celial, and viral growth. The modes of growth serve to high- ing the chemostat a flexible tool to study cellular behavior
light the impact of morphology on bioprocessing consider- under different nutrient limitations. In a turbidostat, the
ations. cell concentration in the vessel is maintained constant by
Bacteria grow by a process of binary fission yielding two monitoring the optical density of the culture and modulat-
identical daughter cells with doubling times typically be- ing the medium feed rate to achieve a set point optical
tween 0.5 and 3.0 h. The high specific growth rates of bac- density. When the optical density rises above the set point,
teria make them especially suitable for many CSTR appli- the feed rate is increased and, because the fluid volume is
cations; however, high O2 requirements and metabolic heat maintained constant by an overflow device, the well-mixed
generation become important considerations in bioprocess culture is diluted and the optical density approaches the
scale-up. Animal cells (5 to 20 lm) are much larger than set point value. The turbidostat is less commonly used be-
bacteria (1 lm) and yeast (5 to 10 lm) cells but grow more cause of difficulties in continuously monitoring the cell con-
BIOREACTORS, CONTINUOUS STIRRED-TANK REACTORS 355

centration. Its main utility is to control the growth rate than as a production technology, that continuous culture
near the maximum growth rate, an operating region in has found its widest and most successful application.
which the chemostat is less stable. A second material balance can be written for the
growth-limiting substrate (S) using an allocation model for
Material Balances. Cell mass is most often used to quan- substrate utilization, in which substrate uptake is divided
tify microbial growth and usually is proportional to cell into cell growth, cell maintenance, and product formation
number under conditions of balanced growth, in which cel- components.
lular chemical and physical properties are preserved in
subsequent generations. Material balances based on cell F • S0 ⳮ F • S
123 123
number may have particular utility for some applications, SUBSTRATE SUBSTRATE
such as mammalian cell culture, where number rather IN OUT

冢 冣
than mass is the conventional method of analysis. One may
qP • X
need to be wary of variations in cell size and morphology l•X
ⳮV• Ⳮ m•X Ⳮ YP /S
that may not be apparent from measurements of cell mass YX /S 123 123
123
or number. The material balance for a uniform cell popu- MAINTENANCE PRODUCT
GROWTH
lation in a CSTR can be written as shown in equation 1, in FORMATION
which l[hⳮ1] represents the specific growth rate and
d(V • S)
␣[hⳮ1] represents the specific rate of cell lysis and/or en-
dt
dogenous metabolism (i.e., resulting in a decrease in cell ⳱ 123 (3)
mass). The specific growth and death rates differ among ACCUMULATION
organisms and are functions of the cell environment (e.g.,
pH, temperature, nutrients). In this equation, YX /S is the cell yield (or dry cell weight,
DCW) on substrate (g DCW/g substrate), m is the cell
d(V • X) maintenance coefficient (g substrate/g DCW ⳯ h), qP is the
F • X0 ⳮ F • X Ⳮ V • l • X ⳮ V • ␣ X ⳱ specific product formation rate (g product/g DCW ⳯ h),
123 123 14243 123 dt
123 and YP /S is the mass yield of product on substrate (g prod-
CELLS CELLS CELL CELL
IN OUT GROWTH LYSIS ACCUMULATION uct/g substrate). Each specific yield coefficient describes
(1) the allocation of substrate to cells, product, or mainte-
nance. Depending on whether the fermentation goal is to
It may be necessary to reformulate the cell balance if produce cells (biomass) or a metabolic product, it may be
the cell population is significantly differentiated; examples possible to simplify the substrate balance by assuming
include mixed cultures, mammalian cell culture, and re- that some uptake terms are dominant. The following sec-
combinant fermentations with plasmid instability (see “Se- tions describe the analysis of biomass production and prod-
lection/Mutation and Contamination”). If we assume the uct formation.
system is operating at steady state and there is no accu- Biomass Production. In the case of biomass production,
mulation of fluid or cells in the vessel, then the time deriv- in which the goal of the fermentation is to produce cells,
ative can be set to zero. Under normal bioprocessing con- the large majority of nutrient uptake goes toward cell
ditions, cell death is assumed to be negligible (i.e., l Ⰷ ␣). growth. The rate of substrate uptake for growth is as-
Bacterial and yeast cells maintain approximately complete sumed to be much greater than that for maintenance (i.e.,
viability except in suboptimal environments and at very l/YX /S Ⰷ m) and product formation (i.e., l/YX /S Ⰷ qP /YP /S).
low dilution rates, whereas in mammalian cell culture, vi- Assuming the system is at steady state, the substrate bal-
ability and cell lysis may vary significantly with time. ance (equation 3) can be rewritten and solved for the cell
Assuming that the feed to the reactor is sterile (X0 ⳱ concentration using equation 2.
O), the material balance reduces to equation 2, in which
the specific growth rate is equal to the liquid flow rate from X ⳱ YX /S • (S0 ⳮ S) (4)
the vessel divided by the liquid volume. This quantity is
called the dilution rate D (the inverse residence time for To determine the relationship between the specific
the CSTR) and has units of hⳮ1. growth rate and the cell environment, a suitable growth
model must be adopted. The simplest and most common
relationship used is the unstructured Monod growth
F model, in which cell growth is a function of a single limiting
l⳱ ⳱D (2)
V substrate, usually the carbon source. Alternative unstruc-
tured growth models are given in Table 1. These unstruc-
Equation 2 illustrates one of the most important attrib- tured growth models are empirically derived from obser-
utes of the chemostat: that the specific growth rate can be vations of chemostat behavior, and their applicability to
controlled by manipulating the dilution rate. Control of the dynamic-batch or fed-batch processes should not be as-
specific growth rate, combined with the ability to maintain sumed. Some unstructured models (e.g., substrate inhibi-
a constant, defined cell environment, makes the chemostat tion) can have more than one solution, making different
a powerful experimental tool with which to investigate the steady states possible depending on the starting condi-
many factors that influence cell growth, metabolism, and tions. Structured models, which typically account for ei-
product formation. It is as an investigative tool, rather ther changes in cell composition, intracellular concentra-
356 BIOREACTORS, CONTINUOUS STIRRED-TANK REACTORS

Table 1. Common Unstructured Growth Models


Monod
lmax • S
l⳱
KS Ⳮ S

Modified monod
lmax • Sk
l⳱
KS Ⳮ Sk

where k is an adjustable parameter

Inhibition models Noncompetitive Competitive


lmax • S lmax • S
l⳱ l⳱
冢 冣 冢 冣
Substrate inhibition S KS
(KS Ⳮ S) • 1 Ⳮ KS Ⳮ S Ⳮ •S
KI KI
lmax • S lmax • S
l⳱ l⳱
冢 冣 冢 冣
Product inhibition P KS
(KS Ⳮ S) • 1 Ⳮ KS Ⳮ S Ⳮ •P
KP KP

tions (4), or cell morphology (5), have been proposed in tion 4, the cell concentration becomes an explicit function
varying degrees of complexity (6,7). Structured models can of D and SO.
have utility when such properties significantly influence
D • KS
冢 冣
the kinetics and are required to accurately describe behav-
ior (such as in dynamic process modeling). X ⳱ YX /S • S0 ⳮ (8)
lmax ⳮ D

lmax • S The biomass productivity of the CSTR (RCSTR), defined as


l⳱ ⳱D (5)
KS Ⳮ S the cell output per reactor volume, is calculated as

F•X D • KS
In the Monod model, lmax is the maximum specific
growth rate of the organism, and KS, called the saturation
RCSTR ⳱
V 冢
⳱ (D • X) ⳱ D • YX /S • S0 ⳮ
lmax ⳮ D 冣
constant, is inversely proportional to the cell’s affinity for (9)
the substrate. The value of KS is typically quite low (1 to 5
mg/L for Escherichia coli on glucose), which means that l Figure 1 shows the steady-state cell and substrate con-
 lmax when S  10KS and that l only becomes a strong centrations and biomass productivity for the ideal chemo-
function of the substrate concentration when S  10KS. stat. Note that the substrate is almost completely utilized
Note that the maximum specific growth rate of the or-
ganism limits the extent to which the dilution rate can be
increased. 16 35
or biomass productivity (g DCW/L × h)

lmax ⳱ DC (6) 14 X
30
Cell concentration (g DCW/L)

This threshold dilution rate is called the critical dilution 12 Substrate concentration (g/L)
25
rate (DC), and increasing the dilution rate beyond DC re-
10
sults in “washout”; cells are removed from the vessel at a 20
rate faster than their growth rate. This can limit the pro- 8
ductivity of the simple CSTR and motivates cell retention
(DX) 15
or recycle strategies that enable operation at higher 6
throughputs.
10
Rearranging the growth model to solve for the substrate 4
concentration in the vessel gives
2 5
D • KS S
S⳱ (7) 0 0
lmax ⳮ D 0 0.2 0.4 0.6 0.8 1
Dilution rate (1/h)
At dilution rates above DC, cells have been washed out of
the vessel, and the substrate concentration equals the inlet Figure 1. Ideal chemostat performance: lmax ⳱ 1.0 hⳮ1, KS ⳱
concentration (SO). By substituting equation 7 into equa- 0.05 g/L, S0 ⳱ 30 g/L, YX /S ⳱ 0.5 g DCW/g substrate.
BIOREACTORS, CONTINUOUS STIRRED-TANK REACTORS 357

over most of the operating dilution rates. This high con- Table 2. Yield Coefficients for Bacteria on Different
version of substrate is a key attribute of the ideal CSTR Carbon Substrates
system, improving the economics of processes that depend Substrate YX /S YX /O2 YHl
on efficient substrate utilization and minimizing the ef- (i) [g DCW/g substrate] [g DCW/g O2] [g DCW/kcal]
fects of substrate inhibition. Also shown in Figure 1 is the
Acetate 0.36 0.70 0.21
high concentration of cells in the vessel until washout at
Glucose 0.51 1.5 0.42
the critical dilution rate. The maximum biomass produc- Methanol 0.40 0.44 0.12
tivity is located close to the washout point, making opera- Ethanol 0.68 0.61 0.18
tion at the maximum productivity point very sensitive to n-Paraffins 1.0 0.50 0.16
deviations in dilution rate. Methane 0.62 0.20 0.061
The dilution rate associated with maximum biomass
Source: From Ref. 16.
productivity for a fixed limiting nutrient concentration in
the feed (DM) can be calculated by setting the derivative of
(DX) with respect to D to zero and solving for the dilution
rate. DM is then given as The biomass productivity of the CSTR can be compared
to the productivity of a batch fermentor (RBATCH) by defin-
ing a relevant batch productivity. The batch fermentation
冢 冪K 冣
KS
DM ⳱ lmax 1 ⳮ (10) cycle consists of a lag phase, an exponential growth phase,
S Ⳮ S0
cell harvest, and a batch turnaround time associated with
cleaning, sterilizing, and filling the vessel. The lag, har-
By substituting DM into equation 8, we can solve for XM,
vest, and turnaround activities can be grouped into a term
the cell concentration corresponding to DM.
tTURNAROUND in order to determine the batch cycle time
(tCYCLE) using equation 13, in which Xi is the concentration
[
XM ⳱ YX /S S0 Ⳮ KS ⳮ 冪KS(S0 Ⳮ KS) ] (11) of cells in the vessel following inoculation (typically Xi 
0.1 ⳯ X).
Considering the limiting nutrient concentration (S0) as

冢 冣
an independent design variable, equation 11 suggests that 1 X
tCYCLE ⳱ ln Ⳮ tTURNAROUND (13)
the substrate concentration in the feed can be increased lmax Xi
arbitrarily to achieve extraordinary cell densities and
productivities. In reality, the productivity of an aerobic re- Cell growth can be calculated from the cell yield on the
actor system is ultimately limited by the rates of heat and/ growth-limiting substrate and the initial concentration of
or mass transfer when the reaction kinetics are fast (i.e., substrate, again assuming cell maintenance and product
high X and D) (see “Mass Transfer”). Oxygen transfer, and formation are negligible.
not the carbon substrate, is often growth limiting because
oxygen is an essential nutrient for aerobic metabolism. It X ⳮ Xi ⳱ YY /SS0 (14)
is poorly soluble in the medium (typically around 7 mg/L
for air at 1 atm), and its transfer rate is restricted by the Subsequently, the ratio of biomass productivities in the
physical capabilities of the oxygenation system. Oxygen- CSTR at DM, (RCSTR)M, and in the batch fermentor is given
limited growth may be expressed as a steady-state balance by equation 15, in which it is assumed that S0 Ⰷ KS (as it
between the oxygen uptake rate (OUR) and the oxygen often is).
transfer rate (OTR). Oxygen transfer is usually limited by
(D • XM)
transfer from the gas to liquid phases, leading to the
(RCSTR)M
⳱ M
X ⳮ Xi
⳱ ln
X
冢 冣
Ⳮ lmax • tTURNAROUD (15)
冢 冣
steady-state balance: RBATCH Xi
tCYCLE

冢 冣
l•X
Ⳮ mO2 • X kLa(C* ⳮ CL) Equation 15 often appears as a measure of relative bio-
YX /O2
1442443 ⳱ 14243 (12) mass productivity, in which the CSTR is favored over batch
OTR
OUR operation at high growth rates and long turnaround times.
However, equations 10 and 11, associated with maximum
in which YX /O2 is the cell yield on oxygen (g DCW/g O2), productivity, do not reflect the ultimate limitation posed
mO2 is the maintenance coefficient for oxygen (g O2 /g DCW by heat and mass transfer in industrial aerobic processes.
⳯ h), kL is the liquid phase mass transfer coefficient (cm/ Given that productivity has a limit dictated by the system,
h) a is the specific interfacial area for mass transfer (cm2 / the independent parameters, D and S0, can be adjusted so
cm3), kLa is the mass transfer coefficient (hⳮ1), and (C* ⳮ that the maximum productivity is attained during opera-
CL) is the driving force for mass transfer where C* is the tion. Recognizing that (RCSTR) is ultimately limited by the
equilibrium oxygen concentration (mmol/L) and CL is the maximum oxygen transfer rate, equation 12 can be re-
dissolved oxygen concentration (mmol/L). Typical values arranged to give the biomass productivity.
for YX /O2 are located in Table 2. The ability of the heat
transfer system to remove heat generated during microbial D • YX /O2 • OTRMAX
(D • X) ⳱ (16)
growth can also limit RCSTR, as discussed in “Energy Bal- (D Ⳮ mO2 • YX /O2)
ance”. These limitations will be important in evaluating
and comparing CSTR performance. Notice that the cell concentration is fixed once an operating
358 BIOREACTORS, CONTINUOUS STIRRED-TANK REACTORS

dilution rate is specified. Solving equation 16 for X and mentation costs (e.g., fuel alcohol or gluconic acid produc-
substituting into equation 4 yields tion), volumetric productivity and conversion yields are es-
pecially relevant criteria as they relate to the size and cost
YX /O2 • OTRMAX of the reactor system and the cost of raw materials. In pro-
S0 ⳱ (17) cesses where recovery costs dominate (e.g., antibiotics),
YX /S • dS • (D Ⳮ mO2 • YX /O2)
however, the size and operational costs of the recovery sys-
where dS is the fractional substrate conversion. High sub- tem are proportional to the fluid volume processed and in-
strate conversion leads to lower raw materials costs and versely proportional to the product concentration (8). As a
reduces the burden of residual substrate on downstream result, the final product concentration, or titer, is more im-
purification and waste treatment operations. When oper- portant than biomass productivity.
ating near the maximum reactor productivity (i.e., maxi- The material balance on the product can be written as
mum OTR), the dilution rate and the inlet substrate con- shown below, in which product formation is expressed us-
centration can be adjusted independently to achieve target ing a specific product formation rate qP (g product/g DCW
levels of cell concentration and/or substrate conversion, as ⳯ h), and product degradation by a specific rate constant
shown in Figure 2. kP(hⳮ1).
For biomass production when O2 transfer is not limiting
productivity, the CSTR is favored over batch operation dP
when the specific growth rate is high and the batch turn- qP • X ⳮ kP • P • D ⳮ D • P ⳱ dt
123 14243 123 123
around time is long. However, most industrial processes PRODUCT PRODUCT PRODUCT PRODUCT
will be operating at or near the oxygen or heat transfer FORMATION DEGRADATION OUT ACCUMULATION
limitation. The desirability of enhanced O2 transfer has (18)
motivated the development of novel bioreactor designs
(e.g., bubble columns, loop or airlift reactors). For thera- Product formation can be characterized in relation to
peutic products, however, the overwhelming majority of growth, being growth (primary metabolites) or nongrowth
fermentations are batch or fed-batch processes. In these (secondary metabolites) associated. Examples of growth-
applications, the choice of operating mode is not based on associated products are direct catabolic products of the car-
biomass productivity. Performance metrics such as volu- bon substrate, such as ethanol and citric or acetic acid.
metric productivity of the product, product yield, and prod- Nongrowth-associated products, comprising many antibi-
uct concentration become more important in evaluating otics, are metabolites that are not necessary for cell growth
potential operating strategies than simply the biomass and typically are only produced during slow or stationary
concentration. growth phases. Some products, such as xanthan gum and
Product Formation. When the goal of the fermentation lactic acid, are mixed growth associated in that they are
is to produce a product other than biomass, the criteria produced during slow and stationary growth phases. Nu-
used to evaluate alternative operational modes are less merous models of product formation have been proposed,
straightforward than biomass productivity. In order to taking into account variables such as hyphal morphology,
evaluate process alternatives, a proper set of performance cell age, surface area, metabolic carbon flux, and plasmid
metrics must be identified that relate to overall process copy number. A simple model expresses the growth depen-
economics. When process economics are dominated by fer- dence of qP as (9)

qP ⳱ ␣•l Ⳮ b
123
100 4 GROWTH NONGROWTH
(19)
Biomass productivity ( g DCW/L*h)
or glucose in feed stream (% w/v)

90 ASSOCIATED ASSOCIATED
Cell concentration [g DCW/L]

3.5
80
(DX) 3 The steady-state product balance can be rewritten and
70
solved for P.
60 2.5
50 2 qP • X (␣ • l Ⳮ b)
冢 冣
b
X P⳱ ⳱ •X⳱ ␣Ⳮ •X (20)
40 1.5 D D D
30
1
20 S0 For growth-associated products (i.e., ␣ Ⰷ b) product con-
10 0.5 centration is proportional to biomass and is independent
0 0 of the dilution rate when X is approximately constant. In-
0 0.02 0.04 0.06 0.08 0.1 creasing the dilution rate results in increased product for-
mation up to the region near DC. The growth dependence
Dilution rate (h–1)
of product formation and intracellular metabolic fluxes can
Figure 2. Chemostat operating at maximum oxygen transfer be determined using a chemostat (10). An example of
rate: OTRmax ⳱ 100 mmol O2 /L ⳯ h, lmax ⳱ 0.09 hⳮ1, YX /O2 ⳱ growth-associated product formation is shown in Figure 3
1.56 g DCW/g O2, mO2 ⳱ 0.024 g O2 /g DCW ⳯ h, YX /S ⳱ 0.45 g (11), in which product concentration is proportional to bio-
DCW/g substrate, 95% substrate conversion. mass and the growth dependence of qP is evident.
BIOREACTORS, CONTINUOUS STIRRED-TANK REACTORS 359

140 6 QAGIT Ⳮ QMET Ⳮ DQSENS ⳮ QLOSS ⳮ QEVAP ⳮ QEXCH ⳱ 0


Total α-amylase activity (units/mL)
(21)

specific α-amylase formation rate


P
120

Cell concentration (OD660) or


X 5
qP

(units/OD660 × mL × h)
in which QAGIT is the mechanical energy imparted to the
100 fluid through impeller agitation (equal to the gassed power
4
input), QMET is the metabolic heat generated by cell
80 X qP growth, DQSENS is the net sensible heat added to the sys-
3 tem by streams entering and leaving the system, QLOSS is
60
the sum of the heat losses from the system to the surround-
2 ings, QEVAP is the latent heat removed by evaporation, and
40
QEXCH is the heat removed from the system by an appro-
20 P 1 priate heat exchanger system (13). In some cases, the heats
of solution and mixing must be accounted for, but in most
0 0 cases they are negligible. The terms DQSENS, QLOSS, and
0 0.1 0.2 0.3 0.4 0.5 0.6 QEVAP are comparatively small, leading to the simplified
Dilution rate (h–1) energy balance

Figure 3. Production of ␣-amylase in a chemostat by recombi- QEXCH ⳱ QMET Ⳮ QAGIT (22)


nant Escherichia coli. The authors (11) used a modified Leude-
king–Piret model, qP ⳱ (␣l Ⳮ b)(1 Ⳮ kl)ⳮ1, to describe ␣-amylase
For fast-growing microorganisms, the heat exchanger duty
kinetics. The term (1 Ⳮ kl)ⳮ1 accounts for an observed increase
can be as high as 7.7 to 23.2 kW/m3, of which QMET and
in plasmid copy number with decreasing growth rate. Model pa-
rameters were regressed from data: ␣ ⳱ 34.12 units/mL ⳯ OD660, QAGIT typically represent about 75% and 25% of the total,
b ⳱ 4.2 ⳯ 10ⳮ10 U/mL ⳯ OD660 ⳯ h, and k ⳱ 8.63 h. Experi- respectively (14).
mental data are depicted as points, model predictions as solid Approximately 40 to 50% of the energy contained by a
lines, and trend lines as dotted lines. substrate is converted into useful chemical energy,
whereas the balance is released as heat. If this metabolic
heat is not removed from the fermentation broth, the tem-
In the case of secondary metabolites (i.e., ␣ Ⰶ b), product perature will rise and possibly hinder performance. Met-
concentration is inversely proportional to the dilution rate, abolic heat generation is a function of the growth rate of
and the productivity (D ⳯ P) is independent of dilution the organism, the cell concentration, the fluid volume, and
rate. The low dilution rates favorable for secondary me- the efficiency of cell growth on a particular substrate (i),
tabolite production approach batch operation, which is which can be expressed as the metabolic heat released per
generally favored over the CSTR in such instances. gram of cell produced (1/YHi) (kcal/g DCW).
Other issues can impact the decision between batch and
continuous culture. The ability of the CSTR to maintain l•X
QMET ⳱ V (23)
an ideal environment for product formation may offer a YHi
competitive advantage over the batch fermentation, with
its time-varying environment and prolonged lag and sta- In aerobic fermentations, oxygen is the final electron ac-
tionary growth phases. Regulatory and market factors also ceptor in substrate metabolism, enabling a correlation be-
play an important role in deciding on the operating mode. tween the rates of oxygen uptake and heat generation. The
The CSTR is a dedicated manufacturing system used to following empirical correlation (15) gives QMET (kcal/h) as
produce a single product. Such a system may not be well a function of specific oxygen consumption, qO2 (mmol O2 /L
suited for the production of specialty chemicals and phar- ⳯ h), and V(L):
maceuticals because it can neither adapt to variable mar-
ket demand nor satisfy demand for multiple products. QMET ⳱ 0.12 • V • qO2 (24)

Energy Balance. Heat transfer is an important consid- where the oxygen demand during exponential growth can
eration in fermentor design, scale-up, operation, and ster- be expressed using the cellular yield on oxygen, YX /O2 (g
ilization. The energy balance is used to determine the DCW/g O2).
time–temperature profile of the fermentation broth by ac-
counting for the transfer and accumulation of energy. The l•X
QO2 ⳱ (25)
heat transfer rate limits the ability to reduce the cycle time YX /O2
for sterilization (12). More importantly, the rate of heat
removal from the broth during cell growth can constrain Table 2 gives values of YHi and YX /O2 for bacterial growth
volumetric productivity, an issue in very large reactors that can be used to estimate oxygen demand and metabolic
with reduced area-to-volume ratios. Because considerable heat generation. Notice that the more reduced substrates
heat generation accompanies rapid cell growth, the high result in greater O2 demand and heat generation and sub-
specific growth rates favored in industrial CSTR applica- sequently a larger burden on the O2 transfer and heat ex-
tions will exacerbate the problem of heat transfer in large change system.
reactors. The steady-state energy balance for the fluid in Heat exchange systems are chosen based on the ex-
the CSTR is written as pected heat exchanger duty, influence on fluid mixing, im-
360 BIOREACTORS, CONTINUOUS STIRRED-TANK REACTORS

pact on cleaning and sterilization, utility economy, and operating supplies, and labor) associated with the CSTR,
maintenance, operating, and capital costs. Heat exchang- it is important to view the reactor system in the context of
ers for fermentors commonly consist of a jacket or shell the entire process. The process flow diagram will form the
around the vessel, internal coolant coils, or occasionally an basis for determining not only the reactor operating costs
external heat exchanger. The rate of heat removal by a but also the capital and operating costs of air compressors,
heat exchanger system can be described using a Fourier’s pumps, holding tanks, and various downstream units. It
law expression: is important to remember that the relative costs associated
with the reactor system, batch or continuous, may not have
QEXCH ⳱ U • A • (T ⳮ Tc) (26) a significant impact on the total process economics if prod-
uct recovery costs are dominant (see “Product Formation”).
in which A is the surface area available for heat transfer, Raw materials costs include the material and handling
U is an overall heat transfer coefficient accounting for all costs of components added to the system to satisfy meta-
heat transfer resistances, and (T ⳮ Tc) is the driving force bolic (e.g., C, N, O sources) or process (e.g., acids or bases
for heat transfer, where T is the bulk temperature of the and antifoam agents) requirements. Direct expenses in-
fermentation broth and TC is the temperature of the cool- clude the cost of utilities, maintenance, operating supplies,
ing fluid used in the heat exchanger. Typical values of the operating labor, direct supervision, laboratory charges,
heat transfer coefficient are 50 to 150 BTU/ft2 ⳯ h ⳯ F and patent royalties. Continuous systems are easier to au-
(280 to 850 W/m2 ⳯ K). An important consequence of fer- tomate and offer the potential of lower labor costs than
mentor scale-up is a decreasing surface-area-to-volume ra- batch production systems, with their labor intensive start-
tio (A/V): at increasing scale the capacity for heat removal up and shut-down operations. Indirect expenses include
relative to heat generation diminishes and often becomes taxes and depreciation, usually expressed as a percentage
limiting at larger volumes. of the plant cost.
The rate of heat removal can be increased by increasing Media. The selection of bioprocess media typically in-
the temperature driving force, increasing the heat transfer volves a trade-off between media cost and the product yield
surface area, or reducing resistance to heat transfer (in- and titer. It is an important stage of development that can
creasing U). Water is primarily used as the coolant fluid influence the design and performance of the entire process.
because of its availability and low cost relative to a refrig- Selecting an apparently cheap media, for example, may
erated coolant system. The temperature of the cooling wa- result in more expensive downstream recovery and waste
ter increases as it passes through the heat exchanger, so treatment operations. In general, media selection includes
an arithmetic or logarithmic mean Tc may give a more rep- a number of technical and economic considerations: yield
resentative measure of the coolant temperature. Increas- or titer of desired and undesired products; cost; variability
ing the coolant flow rate can decrease the mean coolant in composition and price; availability; effect on down-
temperature but with a subsequent increase in utility con- stream processes; need for pretreatment or supplements;
sumption. Furthermore, the pressure drop of the ex- shipment, storage, and handling, and need for testing and
changer and pump capacity limit the extent to which cool- validation.
ant flow rate can be increased. The heat exchanger area is Media can be classified as defined or undefined with re-
dictated by the size and type of heat exchange system cho- spect to chemical composition. Laboratory-scale chemostat
sen during process scale-up. The overall heat transfer co- investigations commonly use defined media to allow pre-
efficient and constituent resistances are thoroughly dis- cise control over the growth-limiting nutrient and medium
cussed elsewhere. In general, poor mixing contributes to composition. Small bioreactors (1 to 4 L) are preferred for
decreased heat transfer (see “Nonideal Mixing”). It is suf- continuous operation in the laboratory because media
ficient to focus on the convective heat transfer coefficient preparation and storage are less of a burden. Industrial
on the fermentation broth side, h[W/m2 * K], which is often microbial fermentations predominantly use undefined me-
the dominant heat transfer resistance. The convective heat dia because they generally are less costly and perform bet-
transfer coefficient is a function of the Reynolds number ter than defined media. Undefined microbial media often
(Re), Prandtl number (Pr), viscosity ratio (Vi), and the sys- contain agricultural by-products (e.g., molasses or corn
tem geometry (FGEOMETRY). In many cases Vi  1, and the steep liquor) and thus are subject to source market fluc-
exponents a and b have typical values of 0.8 and 0.3, re- tuations in quality and price. Processes that have been val-
spectively. idated with a variety of media may change the production
media to take advantage of market changes. The costs of
Nu ⳱ C1 • Rea • Prb • Vic • FGEOMETRY raw materials as a percentage of operating costs for pri-
mary metabolites can range from 40% for citric acid to 70%
a b c
h • DT
冢 冣 冢 k 冣 冢l 冣 F
Di2Nq Cpl l for ethanol from sugar cane (17), whereas for secondary
⳱ C1 • GEOMETRY (27)
k l w metabolites media costs can be around 10 to 20%. At small
scale (less than 10 m3), mammalian cells are often grown
Economic Considerations. Operating costs are those in undefined serum, an expensive and sometimes scarce
costs associated with maintaining a given production level media derived from mammalian plasma. Viral contami-
dictated by the scale and scheduling of the process, typi- nation and the presence of serum proteins can complicate
cally grouped as raw materials, direct expenses, and in- cultivation, product recovery, and quality control and qual-
direct expenses. Although the focus here will be on raw ity assurance (QA–QC) in industrial processes. These com-
materials and direct expenses (e.g., utilities, maintenance, plications, combined with regulatory pressure to safeguard
BIOREACTORS, CONTINUOUS STIRRED-TANK REACTORS 361

against viral contamination, are a driving force toward the are critical to mixing. High broth viscosities, typical of my-
development of defined, serum-free media in order to avoid celial fermentations, contribute to poor mixing. Poor mix-
use of animal-derived media components. ing can affect oxygen transfer (18), product formation (19),
Utility Expenses. Steam, cooling water, and power re- heat transfer, process monitoring and control, and the dis-
quirements comprise the majority of utility expenses. Pro- tribution of components added to the system (20,21).
cess demand for water-for-injection (WFI) must also be de- Nonideal mixing in continuous systems is typically
termined for clean-in-place (CIP) systems. Steam usage characterized by either the residence time distribution, the
occurs predominantly during media and equipment ster- distribution of fluid residence times around the ideal res-
ilization and can be calculated using knowledge of the ster- idence time (s) (22), or the mixing time (tM), the time re-
ilization cycle. Continuous reactors readily lend them- quired for a system to respond to a feed disturbance. The
selves to the use of continuous, as opposed to batch, media ideal mixing time is equal to zero (instantaneous), and the
sterilization systems, offering advantages in reduced ther- mean value of the residence time distribution is s (or Dⳮ1).
mal degradation of heat-sensitive media, reduced sterili- Residence time distributions are typically measured by
zation time, more efficient fermentor use, and greater pulse or step addition of a tracer into the reactor feed and
steam economy (about 20 to 25% of the steam used by tracking the appearance of the tracer in the exit stream as
batch sterilization) (12). The cooling water requirement, a function of time. Correlations obtained from dimensional
wH2O (kg/h), can be calculated knowing the heat exchanger analysis can be used to predict mixing times, which in-
duty from the energy balance (equation 21): crease with scale and broth viscosity (23). Small vessels
(500 L) are generally well mixed (tM  s), but large fer-
QEXCH mentors (5,000 L) typically have poor mixing (tM  min).
wH2O ⳱ (28)
Cp • (Tout ⳮ Tin) Compartmental mixing models, in which the bulk fluid is
modeled as discrete CSTRs and plug flow reactors (PFRs)
where Cp is the heat capacity of water [1 kcal/kg * K] and with fluid interchange, have been used to describe hetero-
Tout and Tin are the outlet and inlet cooling water tem- geneity in large vessels (24).
peratures, respectively. The availability of abundant low- Wall growth is a special case of nonideal mixing in
temperature water can reduce cooling water requirements which cells adhere and proliferate on vessel surfaces (25),
and possibly allow the use of larger reactors because of the where they are hidden from cell mass measurements ob-
improved ability to remove metabolic heat. tained from samples of the bulk fluid. The system, in ad-
Oxygen demand and mixing requirements drive power dition to being heterogeneous, is no longer at steady state
consumption by agitators and compressors, whereas larger because cells are accumulating in the vessel. This can pose
volumes, higher cell densities, higher specific O2 uptake
serious problems if the accumulating organism is a con-
rates, and higher broth viscosities result in increased
taminant or otherwise undesirable organism. The metab-
power requirements. Total power consumption for agitated
olism and growth of wall-bound cells can be quite different
vessels is typically in the range of 2 to 10 kW/m3. Centrif-
from the suspended population because of mass transfer
ugal pumps are predominantly used in bioprocesses for
limitations. Wall growth can reduce heat transfer, create
their relatively low cost and ability to handle suspended
sterilization and cleaning problems, and corrupt measure-
solids. Although pumps represent only a small fraction of
ments in experimental systems. It can be a significant fac-
power consumption, they can be a significant percentage
tor when the surface-area-to-volume ratio (A/V) is high
of overall maintenance costs. In addition, the performance
(e.g., laboratory-scale systems and vessels with internal
of continuous processes is particularly susceptible to pump
cooling coils) and may require modifications to chemostat
failure and may warrant additional capital investment to-
ward the installation of backup pumps in parallel. design or operation (26). For example, the glass walls of
laboratory vessels are sometimes treated with organosi-
Actual vs. Ideal Behavior. In this section, the assump- lane compounds to minimize wall growth. Yeast and my-
tions used in the development of the ideal CSTR theory are celial cells with a propensity to form pellets are prone to
revisited in order to determine when they are invalid and wall growth under the same conditions that favor floccu-
to gauge the impact of nonidealities on performance. lation. A discrete washout point does not exist with wall
Nonideal Mixing. The major assumption of the ideal growth, because cells are effectively immobilized in the
CSTR is that there are no spatial variations of properties vessel even past DC.
inside the vessel. Such ideal mixing is never observed in Substrate Assumptions. Deviations from ideal chemostat
actual systems and even deteriorates on scale-up, although behavior may arise when assumptions regarding the mag-
it can be a good approximation of behavior. The challenge nitudes of nutrient uptake, the consistency of cell compo-
is to determine when nonidealities can be expected and sition, or the identity of the growth-limiting nutrient be-
how they will influence performance. come invalidated. Substrate uptake for growth in equation
Equipment design, operating conditions, and broth 3 was assumed to be much larger than that for mainte-
properties all influence the quality of fluid mixing in the nance and product formation. At low dilution rates (i.e.,
vessel. Agitated tanks are designed to provide good mixing low growth rates) the maintenance term becomes signifi-
through selection of tank geometry, baffle placement, and cant (m  l/YX /S) and less substrate goes toward cell
impeller design, although mixing quality decreases with growth, causing the actual X to be less than predicted by
increasing scale. Agitator power input and gas sparging the ideal theory at low D. Similarly, at high dilution rates,
rates, although typically associated with oxygen transfer, the production and accumulation of growth-related prod-
362 BIOREACTORS, CONTINUOUS STIRRED-TANK REACTORS

ucts and intermediates can become significant, leading to through the application of elevated temperatures, ex-
a reduced cell yield at higher D. tremes of pH, the use of narrowly defined or modified me-
The ideal chemostat derivation assumes that cell com- dia, and the use of specially selected cultures (e.g., anti-
position does not change over the operating region. In ac- biotic-resistant strains).
tuality, cell composition varies with pH, temperature, The prolonged operating periods of continuous culture
growth rate (27), and medium composition. As cell com- increase the probability of contamination by a foreign or-
position changes, the demand for essential nutrients will ganism. The threat posed by contamination depends on the
change in ways that were not accounted for in the derived ability of the undesirable microorganism to complete and
equations. Proper evaluation of experimental data from a thrive in the CSTR environment. Consider the case of two
chemostat may require consideration of the variation of types of microorganisms with concentrations X (the de-
nutrient uptake and cell composition with environment sired strain) and Z (the contaminant) competing for the
and growth rate. same limiting substrate in a CSTR. The material balances
When using complex and undefined media or an organ- on cell mass, neglecting cell lysis, can be written as
ism with complex nutritional requirements (e.g., mam-
malian cells), it is often difficult to identify the growth- dX
⳱l•XⳮD•X (29)
limiting nutrient. In addition, the limiting nutrient itself dt
may change because the nutrient demand, and subse-
dZ
quently the media composition, may change with operat- ⳱ lz • Z ⳮ D • Z (30)
ing conditions. The cell concentration profile in these sit- dt
uations likely would be constantly decreasing with dilution
rate, unlike the ideal chemostat where X is approximately Subtracting equation 30 from equation 29 and rearranging
constant over the majority of 0  D  DC. More compli- leads to equation 31:
cated, structured growth models would be required to ac-
d ln[X/Z]
count for such behavior. ⳱ l ⳮ lz (31)
Non-Steady-State Behavior. The potential sources of pro- dt
cess variability and non-steady-state behavior are perhaps
too numerous to mention. However, typical instances in- which shows that the growth rates and their dependence
clude chemostat start-up, execution of control actions, in- on the limiting substrate will determine the fate of the
duced disturbances (e.g., pulse and shift methods [28]), culture (32). The contaminant (Z) could be washed out (l
variations in feed composition, culture degeneration (e.g.,  lz), remain at a stable level (l ⳱ lz), or dominate (l 
plasmid loss or apoptosis), wall growth, or equipment fail- lz) the culture. This simple analysis of selection can be
ures. Steady state often is declared when the measurable complicated if the contaminant has properties that prevent
process states are maintained constant for 3 to 5 residence it from being washed out (e.g., adhesion to reactor sur-
times. Sustained oscillations are sometimes observed in faces), if the contaminant competes for a different sub-
continuous culture and often are the result of growth in- strate than the production strain, or if there are interac-
hibition resulting from either an accumulated product (29) tions from inhibitory cellular products. For example,
or the burden of product formation (30). lactobacilli are often a persistent contaminant of continu-
Selection/Mutation and Contamination. By controlling ous ethanol production processes because of their intimate
the culture conditions in the CSTR, a highly selective en- association with flocculant yeast aggregates and ability to
vironment for the selection and proliferation of certain mi- adapt to high ethanol concentrations (33). Selective recycle
croorganisms can be created. Cells in this selective envi- of desirable organisms back to the vessel has been used to
ronment with growth rates less than the dilution rate will prevent domination of the culture by undesirable strains
be washed out of the reactor, leaving only those cells with (34–36).
the properties that have been selected for. In this way, con-
tinuous culture can be used as a strain improvement tool Enzymes
to select organisms that possess a desirable trait, such as Introduction. Enzymes are biological catalysts with
yeast with higher ethanol tolerance (31). high selectivity toward reactants and products, making
Because of the metabolic burden imposed by high levels them attractive for use in a number of industrial applica-
of product formation, the production strain has a growth tions. Enzyme activity is strongly influenced by the envi-
disadvantage relative to unproductive strains that are ronment (e.g., pH, temperature, metal ions). Loss of activ-
present in the reactor. Without selection pressure in favor ity or denaturation can be reversible or irreversible,
of the production strain, a gradual decline in productivity depending on the type, strength, and duration of an un-
will be observed over time as nonproductive cells dominate favorable interaction. A benefit of using the CSTR for en-
the culture. Examples include the reversion of specially zyme reactions is that the constant, controlled reactor en-
selected antibiotic strains to low productivity mutants or vironment can be designed for maximum enzyme activity
the domination of recombinant protein processes by and life.
plasmid-free cells. The configuration of CSTRs in series
can be used to circumvent this problem by providing sepa- Material Balances. Assuming that the inlet and outlet
rate environments for growth and product formation (see flow rates are approximately equal (i.e., solutions are di-
“CSTRs in Series”). Selective pressure in favor of the pro- lute), the steady-state material balance on the substrate
duction strain (and against contaminants) may be exerted can be written as
BIOREACTORS, CONTINUOUS STIRRED-TANK REACTORS 363

F • (S0 ⳮ S) ⳱ v • V (32) actions. Higher temperatures also result in increased rates


of thermal denaturation and loss of the active biocatalyst.
where F is the volumetric flow rate, v is the rate of sub- Process economics often depend on optimal temperature
strate consumption by reaction, V is the fluid volume in control to maintain high substrate conversion and long cat-
the reactor, and S0 and S are the substrate concentrations alyst life (38–40). The effects of temperature on the cata-
in the feed and vessel, respectively. Rewriting equation 32 lytic rate constant (k2) can be described using an Arrhenius
in terms of the fluid residence time (s) and the fractional expression
substrate conversion (dS) yields the CSTR design equation
k2 ⳱ A • eⳮEa /RT (34)
V S •d
s⳱ ⳱ 0 S (33) where A is the Arrhenius constant, Ea is the activation
F v
energy, R is the gas constant, and T is the absolute tem-
Using a valid rate expression for v, the design equation perature. The activation energy of enzyme-catalyzed re-
can be used to determine the reactor volume required to actions ranges from 4 to 20 kcal/mol, with most reactions
yield a given conversion rate (S0 ⳯ dS ⳯ F). Fast reaction near 11 kcal/mol.
kinetics are obviously favorable because reactor cost scales Thermal denaturation usually can be described as a
with reactor size. Enzyme loading in the reactor can be first-order decay reaction:
increased beyond the solubility limit by immobilization on
inert support particles, which increases v, reduces the nec- dE
⳱ ⳮkd • E or E ⳱ E0eⳮkdt (35)
essary reactor volume, facilitates enzyme retention and re- dt
cycle, and may improve enzyme stability.
Enzyme Reaction Kinetics. Numerous mechanistic mod- where kd is the thermal denaturation constant, which also
els have been developed to describe enzyme reaction rates follows an Arrhenius temperature dependence. For ther-
as a function of enzyme and substrate concentrations. mal denaturation, Ea varies from 40 to 130 kcal/mol, with
Some of the more common models appear in Table 3 ac- most in the vicinity of 70 kcal/mol. Increasing temperature
companied by the corresponding solution to the design has a greater effect on the rate of denaturation than ca-
equation (equation 33). talysis. For a typical enzyme (i.e., 11 and 70 kcal/mol), an
Unlike a plug flow reactor (PFR), in which the substrate increase in temperature from 30 to 40 C results in a 1.8-
enters at a high concentration and leaves at a lower con- fold increase in the rate of catalysis, but a 41-fold increase
centration, the substrate concentration in a CSTR is at a in the denaturation rate.
uniform, low concentration. The reduced substrate concen-
tration leads to a slower reaction rate, so that the CSTR Energy Balance. An optimal temperature control strat-
requires more of the active enzyme than the PFR to attain egy requires good heat removal because most industrial
the same substrate conversion rate. Substrate inhibition enzyme reactions are exothermic. Although heat transfer
is less problematic in a CSTR than in a PFR because of the is generally good for soluble enzymes in agitated tanks, the
lower substrate concentration in the bulk fluid, whereas high enzyme concentrations attained with immobilization
product inhibition is generally more of a problem in CSTRs can result in fast reaction rates and appreciable heat gen-
than PFRs. Arranging CSTRs in series can reduce the ef- eration. Heat transfer resistance within the catalyst pellet
fects of product inhibition (approaching PFR behavior) can reduce heat removal rates, resulting in higher pellet
while taking advantage of the good mixing characteristics temperatures and shorter catalyst life. Catalyst degrada-
of the CSTR to provide optimal pH control (37). tion from insufficient heat removal is more of a concern in
Temperature Effects. Like many chemical reactions, in- packed beds with high enzyme loading, where heat trans-
creasing the temperature enhances the rate of enzyme re- fer resistances can be significant. The steady-state energy

Table 3. Some Common Enzyme Kinetic Expression


Rate expression Design equation
v⳱ s⳱
vmax • S
冤 冥
1 dS
Michaelis–Menten S0 • dS Ⳮ Km •
Km Ⳮ S vmax 1 ⳮ dS
vmax • S
冤S 冥
1 dS S02
Substrate inhibition S2 0 • dS Ⳮ K⬘m • Ⳮ • (dS ⳮ dS2 )
S Ⳮ K⬘m Ⳮ vmax 1 ⳮ dS K⬘S
K⬘S
vmax • S
K⬘m S0 • dS2
冤S 冢 冣冥
1 dS
• dS Ⳮ K⬘m • Ⳮ •
冢 冣
Competitive product inhibition P 0
K⬘m 1 Ⳮ ⳭS vmax 1 ⳮ dS KP 1 ⳮ dS
KP
Where vmax ⳱ k2 ⳯ E0 Where the product concentration (P) is related to converted substrate;
P ⳱ dS ⳯ S0
364 BIOREACTORS, CONTINUOUS STIRRED-TANK REACTORS

balance for an exothermic (DHRXN  0) enzyme reaction in tion (cO2) of 7 mg/L, and the following cell parameters (42):
a well-mixed CSTR is l ⳱ 0.075 hⳮ1, YX /O2 ⳱ 1.56 g DCW/g O2, and mO2 ⳱ 0.024
g O2 /(g DCW ⳯ h). The Damkholer number shows that
FqcP(TF ⳮ T) Ⳮ v⬘ • E • V(ⳮDHRXN) ⳮ QEXCH ⳱ 0 (36) internal mass resistance is considerable and that cell
growth in the pellet is likely to be limited by oxygen trans-
where F is the volumetric flowrate (m3 /h), q is the fluid fer.
density (kg/m3), cP is the fluid heat capacity (kJ/kg ⳯ K),
v⬘ is the specific reaction rate (kmol substrate/kg enzyme PELLET
O DEMAND VOLUME
⳯ h), E is the enzyme concentration (kg enzyme/m3), 64748
2 PELLET
CELL
678
DHRXN is the heat of reaction (kJ/kmol substrate), QEXCH DENSITY
冢Y 冣 冢3冣
l 4pr3
is the heat removed by the heat exchanger (kJ/h) from Ⳮ mO2 • • •
X /O2 X
equation 26, and TF and T are the feed and bulk fluid tem- Da ⳱ ⳱

冢 冣
cO • (4 • p • r2)
peratures (K), respectively. DO2 • 2 123
r
123 PELLET
Economic Considerations. For enzyme CSTRs, the pri- DIFFUSIVE FLUX SURFACE AREA
mary operating costs are associated with enzyme replace- 0.075 hⳮ1
冢1.56 g DCW/g O 冣
0.024 g O2 0.1 g DCW
ment. Prolonged catalyst activity and marked reductions Ⳮ • • (4 ⳯ 10ⳮ2 cm)2
2 g DCW • h cm3
in raw materials costs can be achieved by maintaining an ⳮ6
3 • (1 ⳯ 10 cm /s) • (7 ⳯ 10ⳮ6 g/cm3)
2
optimal environment for the enzyme during operation.
艑 5.5 ⳯ 105 (38)
Preserving enzyme activity reduces the number and fre-
quency of labor-intensive cleaning and changeovers, facili-
Gas–liquid mass transfer is often rate-limiting for gases
tating downstream operations by consistently providing a
that are sparingly soluble in the broth, such as oxygen and
constant-quality product stream. With multiple reactors
methane. Although highly soluble, carbon dioxide exhibits
installed in parallel, changeovers can be scheduled to min-
pH-dependent partitioning between gaseous and dissolved
imize production variations and downtime. Industrial en-
forms (CO2, H2CO3, HCOⳮ 2ⳮ
3 , CO3 ) that is influenced by the
zymes are often sold as a crude mixture containing only a
rates of both reaction and mass transfer. Proper interpre-
fraction of active enzyme. Selecting enzymes among differ-
tation of the respiratory coefficient (RQ) in fermentations
ent vendors may involve a trade-off between cost and pu-
operated at neutral pH requires consideration of CO2 dy-
rity (percent active enzyme) and a consideration of how the
namics. The OTR has already been used to determine the
impurities may affect the process.
productivity limit of a CSTR used for biomass production
in equation 12. In general, the rate of mass transfer from
MASS TRANSFER the gas to the liquid phase is given as
Introduction NA ⳱ kLa(C* ⳮ CL) (39)
The rate of mass transfer ultimately will limit the maxi-
mum aerobic reactor performance. Oxygen transfer to the where NA is the rate of gas transfer (mmol/l ⳯ h) and the
fermentor broth, for example, can limit both the extent and remaining terms have the same definitions as in equation
rate of cell growth. Mass transfer limitations to microbial 12. For sparged, agitated tanks, kLa has typical values in
flocculants and immobilized catalyst pellets can result in the range from 50 to 1,400 hⳮ1. Several correlations have
reduced reaction rates and inefficient conversion. Liquid– been developed for kLa as a function of the gassed power
liquid mass transfer rates from hydrocarbon substrates to input per unit volume and the superficial gas velocity for
suspended cells may limit productivity in two-phase sys- Newtonian broths in a variety of fermentors (43). The cor-
tems (41). To determine the rate-controlling regime, it is relations can offer wide variability in mass transfer esti-
useful to characterize the relative rates of mass transfer mates and should be used in conjunction with knowledge
and reaction using the dimensionless Damkohler number from past experience or empirical measurements of kLa
(Da): (e.g., dynamic or sulfite oxidation methods). Oxygen trans-
fer to shear-sensitive mammalian cells requires gentle ag-
Maximum rate of reaction itation combined with surface or membrane aeration, or
Da ⳱ (37)
Maximum rate of diffusion light sparging, as opposed to the large power inputs and
high rates of gas sparging in microbial fermentations.
The observed reaction rate may be limited by the rate of This limitation is somewhat offset by the fact that mam-
diffusion depending on the value of the Damkohler num- malian cells have lower O2 requirements (0.05 to 0.5 mmol/
ber: if Da Ⰷ 1 the diffusion rate is limiting, if Da Ⰶ 1 the 109 cells ⳯ h) (44) and grow to lower cell densities (106 to
reaction rate is limiting, and if Da  1 then the reaction 107 cells/mL) than microbial cultures.
and diffusion rates are comparable. As with all dimension-
less numbers, the Damkohler number is only meaningful
VARIATIONS ON THE SINGLE CSTR
if it is calculated using the proper time and length scales
for a given system. Consider spherical pellets (r ⳱ 400 lm)
Single CSTR with Recycle
of Penicillium chrysogenum, assuming a pellet cell density
eff
of 0.1 g/cm3, an effective oxygen diffusivity (DO2 ) of 1 ⳯ Volumetric productivity is related to the concentration of
ⳮ6 2
10 cm /s, a particle size of 400 lm, an oxygen concentra- active catalyst. Cell or enzyme concentrations greater than
BIOREACTORS, CONTINUOUS STIRRED-TANK REACTORS 365

the steady state obtained from the simple CSTR can be fluid density, g is the gravitational acceleration constant,
achieved by separating cells from the effluent stream and and g is the fluid viscosity. The functional dependence of
recycling them to the vessel (27,45) or by retaining them Stoke’s law suggests ways of increasing the settling veloc-
within reactor. Higher catalyst concentrations enhance ity. The easiest and most common method is to increase
substrate conversion and reduce the reactor size necessary the effective cell size by promoting flocculation (cell aggre-
to attain a given conversion. Recycle operation improves gation) through physiological, chemical, and physical fac-
system stability in the face of feed disturbances by retain- tors: selection of flocculant strains; modification of cell wall
ing cells in the vessel even under conditions that would structure or surface charge; changing the pH, tempera-
cause washout in the simple CSTR. Recycle systems can ture, or shear stress; addition of inorganic salts (e.g., Ca2Ⳮ
be operated at dilution rates, or throughputs, greater than and Mg2Ⳮ salts) or clays; controlling the concentration of
the specific growth rate of the organism. Productivity im- certain nutrients or products (e.g., extracellular polysac-
provements achieved with cell recycle are demonstrated in charides); and controlling the cell age or growth phase. Se-
Table 4 for Saccharomyces cerevisiae ATCC 4126 and Zym- lective cell recycle has been implemented using the differ-
omonas mobilis ATCC 10988 at 100 g/L glucose feed. ential sedimentation properties of a desired and unwanted
microorganism (34–36). The properties of a particular
Cell Recycle Methods. Cell recycle is implemented broth are generally unchangeable and will probably only
through a cell separation step, often by a unit operation impede particle settling.
commonly used in the initial stages of downstream pro- Although equation 40 holds for dilute suspensions of
cessing. Typical methods of continuous cell separation in- cells, the interactions among settling particles in concen-
clude centrifugation, filtration, and sedimentation. Cell trated slurries results in hindered settling. The hindered
separation can be viewed as having two often equally im- particle velocity (uh) is influenced by the particle concen-
portant purposes: (1) the recovery or retention of cells for tration and can be expressed with the following correlation
reuse and (2) the removal of potentially inhibitory by- (49):
products or products from the culture environment. The
separation step often has to satisfy additional performance uh 1
⳱ (41)
requirements such as handling of shear- or temperature- u0 1 Ⳮ k • ⑀1P/3
sensitive materials, selectivity in rejection or recovery, con-
tainment, maintenance of asepsis, corrosion resistance, in which ⑀P is the volume fraction of particles and k is an
brief retention time, and ease of cleaning, sterilization, empirical function of ⑀P. For dilute suspensions, ⑀P  0.15,
maintenance, and validation. Recycle operation is stan- whereas in slurries 0.15  ⑀P  0.50.
dard for reactors using stable enzymes, because discarding The limiting settling velocity for a system has a strong
expensive active catalyst is economically unfeasible. Cell influence on equipment design and operation. Consider the
separation operations are discussed elsewhere in the con- case of a continuous sedimentation tank with volumetric
text of downstream processing, although a brief descrip- throughput (F) and constant cross-sectional area (A). The
tion is presented here in relation to cell recycle. sedimentation tank performance can then be described by
Sedimentation. Sedimentation is the settling of parti- equation 42, in which throughput is directly proportional
cles in a gravitational field. With low energy requirements to A and independent of tank depth:
and simple equipment, sedimentation is a relatively inex-
pensive way of separating a dilute cell phase. Waste treat- F ⳱ ulim • A (42)
ment is by far the largest application of sedimentation-
based cell recycle, in which cells are typically separated in The throughput and limiting settling time will thus dictate
large sedimentation tanks using lime or clay to enhance equipment size and costs. Another design consideration is
flocculant formation. The settling velocity (u0) for an iso- the residence time of cells in the settling tank, which must
lated spherical particle can be described using Stoke’s law: be considered in the context of nutrient depletion (particu-
larly for oxygen) and its potential effects on performance.
d2P(qP ⳮ qF)g Sedimentation at laboratory scale may be implemented
u0 ⳱ (40)
18g with an external settling column (47,50,51). Similar de-
vices may be used at bioprocessing scales, whereas large
in which dP is the particle diameter, qP is the particle den- open-air tanks must be used in high-volume wastewater
sity (the specific gravity of a typical cell is 1.05), qF is the treatment. Internal sedimentation has been implemented

Table 4. Ethanol Productivity Enhancements for S. cerevisiae and Z. mobilis


Dilution rate Ethanol productivity Cell density
(hⳮ1) g/(L ⳯ h) (g DCW/L) Reference
S. cerevisiae CSTR 0.17 7 12 46
CSTR with recycle 0.68 29 50 47
Z. mobilis CSTR 0.175 8 2.5 29
CSTR with recycle 2.7 120 38 48
366 BIOREACTORS, CONTINUOUS STIRRED-TANK REACTORS

in tower fermenters, in which immobilized cells and en- membranes have the potential for complete cell recycle, a
zymes or microbial flocculants are retained in the vessel purge or bleed stream is typically split from the recycle
by a sedimentation zone within the vessel. Unlike the ideal stream to prevent accumulation of inert particles and de-
CSTR, tower fermenters may exhibit spatial variations in bris in the vessel.
nutrient concentrations and broth properties along the
height of the tower that can significantly influence reactor Material Balances. A schematic of a CSTR with recycle
performance. In addition, the productivity in these reac- of cells is shown in Figure 4. A material balance on cell
tors may be limited by the need to maintain low upward mass for the CSTR with recycle system, neglecting cell
velocities (e.g., low aeration or CO2 evolution) to allow ad- death, may be written
equate cell sedimentation.
Centrifugation. The operating principle behind centri- F • X0 Ⳮ ␣ • F • C • X1 ⳮ (1 Ⳮ ␣) • F • X1
fugation is the same as that of sedimentation; however, 123 14243 1442443
much higher settling velocities than in sedimentation may CELLS IN CELLS IN CELLS OUT
FEED RECYCLE STREAM
be obtained in the centrifugal field. Centrifugal separators
d(V • X1)
enable high-volume continuous processing of fluids con- Ⳮ V • l • X1 ⳱
123 dt
taining many particles, with short retention times and 123 (44)
CELL GROWTH
small space requirements. To determine the unhindered CELL ACCUMULATION
particle velocity in a centrifugal field (u0C), equation 40 is
multiplied by the centrifugal coefficient (C), also known as where ␣ is the recycle ratio equal to the recycle volume
the G-value, which describes the increase in sedimentation divided by the feed volume, C is the concentration factor
rate due to centrifugation relative to gravitational settling: (cell concentration in the recycle divided by the effluent
concentration) related to the efficiency of the separation

ⳮq ) 冢 g 冣
rx2 step, and X0, X1, and X2 are the cell concentrations in the
d2Pg(qP F feed, recycle, and separator effluent streams, respectively.
u0C ⳱ 123 (43) Note that the low substrate concentrations in waste treat-
18g
C ment create suboptimal growth environments in which cell
death cannot be neglected.
where r is the radial distance from the axis of rotation and Assuming the system is at steady state (dX1 /dt ⳱ 0) and
x is the angular velocity. that the feed is sterile (X0 ⳱ 0), equation 44 yields
Industrial centrifuges are most often classified by in-
ternal structure (e.g., disk stack, tubular bowl) and mode l ⳱ (1 Ⳮ ␣ ⳮ ␣ • C) • D (45)
of operation (e.g., solids retaining, continuous or intermit-
tent solids ejecting). The selection of sturdier construction The dilution rate is no longer equal to the specific growth
and materials will enable higher rotation speeds for sep- rate; in fact, because C  1 and ␣  1, the dilution rate is
aration of smaller particles. The equation describing greater than the specific growth rate.
throughput in a centrifuge is analogous to equation 42, ex- A material balance on the limiting substrate, again
cept that the centrifuge area is expressed using the R neglecting maintenance and product formation, may be
value, which is the area equivalent for a given centrifuge written
and rotation speed. Centrifuge manufacturers will often
provide machine-specific R values, although the R value
for simple disk-stack and tubular-bowl centrifuges can be
calculated directly. F
Filtration. Filtration is separation based on size, allow- S0
ing retention of molecules larger than the pore size of the (1 + α) F
filter and passage of smaller molecules. Membrane filtra- X1
tion thus offers the twin benefits of cell retention and in-
hibitory by-product removal. In cell recycle systems, the
most common arrangements are internal filters for cell and
enzyme retention (52,53) or external membrane filters
(54,55) in plate and frame, spiral cartridge, and hollow fi-
ber configurations. In all these configurations, flow pat- Cell F
terns tangential to the membrane surface can reduce foul- separator
X2
ing and improve the filtrate flux across the membrane.
Compared to internal filters, external filters have higher Volume V
surface-area-to-volume ratios and may be easier to main-
tain; however, they may be less easily sterilized (particu-
larly for some polymer membranes) and could introduce αF
problems of nutrient depletion in the external recycle loop. C*X1
Membrane selection depends primarily on the critical par-
ticle size, with other criteria being cost, mechanical stabil- Figure 4. CSTR with recycle. Cell separation can be achieved
ity, and susceptibility to plugging and fouling. Because through centrifugation, sedimentation, or filtration.
BIOREACTORS, CONTINUOUS STIRRED-TANK REACTORS 367

F • S0 Ⳮ ␣ • F • S ⳮ (1 Ⳮ ␣) • F • S higher biomass productivity of the recycle system results


123 123 1442443
from a dilution rate higher than the specific growth rate
SUBSTRATE SUBSTRATE SUBSTRATE
IN FEED IN RECYCLE OUT
and the increased cell concentration in the vessel.
l•X dS Implementation of a recycle system is often critical to
ⳮ Y •V ⳱ V• the economic viability of processes using expensive biocat-
dt
123
X /S 123 (46) alysts (e.g., enzymes). Typically, this is accomplished by
SUBSTRATE SUBSTRATE immobilizing the enzymes on inert support particles to fa-
CONSUMED ACCUMULATION
cilitate either internal or external recycle. The potential
use of cell recycle in an industrial process involves weigh-
Solving equation 46 for the cell concentration, assuming ing the effectiveness and economics associated with the cell
steady-state (dS/dt ⳱ 0) operation, yields separation step against the marginal improvement in pro-
cess performance. It should also be noted that higher cell
YX /S • (S0 ⳮ S) densities exacerbate the oxygen transfer and heat removal
X1 ⳱ (47)
[1 Ⳮ ␣(1 ⳮ C)] burden of the system.

in which the steady-state cell concentration with recycle is CSTRs in Series


greater than that in the simple CSTR by a factor of 1/[1 Ⳮ In the single CSTR, the constant, controlled environment
␣(1 ⳮ C)]. By adopting a suitable expression for cell gives the advantage of being able to control the cellular–
growth, the substrate concentration can be determined. enzyme environment for maximum utility. Sometimes,
Using the Monod expression, as before, and solving for the however, a particular cell system will exhibit multiple
substrate concentration gives properties of interest that can only be realized in different
environments. The optimal environments for cell growth
l D(1 Ⳮ ␣ ⳮ ␣C) and product formation, for example, may be characterized
S ⳱ KS ⳱ KS (48)
lmax ⳮ l lmax ⳮ D(1 Ⳮ ␣ ⳮ ␣C) by different temperatures, pH, and limiting nutrients. The
configuration of CSTRs in series lends itself to those ap-
Substituting equation 48 into equation 47 yields plications in which multiple environments are required.

D(1 Ⳮ ␣ ⳮ ␣C) Cell Growth. Bacterial growth in the presence of mul-


X1 ⳱
YX /S

S ⳮ KS
(1 Ⳮ ␣ ⳮ ␣C) 0 lmax ⳮ D(1 Ⳮ ␣ ⳮ ␣C) 冥 tiple carbon substrates often results in diauxic growth, in
which cells preferentially metabolize a single substrate
(49)
over all others. In a waste treatment application, the pre-
ferred substrate would be consumed by the microorgan-
A material balance on cell mass around the separator gives
isms and the remaining substrates would pass through the
the cell concentration in the outlet:
system untreated. Configuring CSTRs in series provides a
partitioning of cell metabolism so that less-favored sub-
X2 ⳱ (1 Ⳮ ␣ ⳮ ␣ • C) • X1 (50) strates are consumed in subsequent stages.
CSTRs in series have been used to improve recombinant
Figure 5 shows the cell mass and biomass productivity of protein fermentations in which performance is threatened
a CSTR with recycle compared to a simple CSTR. The by plasmid instability (56,57) and lethal protein overpro-
duction (30). Cells are grown to high density in the first
stage without inducer so that plasmid-free cells have little
35 30 growth advantage over plasmid-containing cells. Induction
Biomass productivity [g DCW/L × h]

in the second stage results in higher productivity than the


Cell concentration [g DCW/L]

30 25 simple CSTR because the continuous introduction of


X
plasmid-containing cells from the first stage reduces the
25
(DX2) 20 ability of nonproductive cells to dominate the culture.
Consider the two-stage system for biomass production
20
in Figure 6, in which a separate feedstream can be added
15
15 to the second stage. The steady-state material balances for
X X2 cells and substrate in the first reactor are identical to the
10 single CSTR case (equations 1 and 3), with steady-state
10
solutions as equations 5 and 8 for Monod growth. Consid-
DX 5 ering the case without the second feedstream, the material
5
balances on cell mass and growth-limiting substrate in the
0 0 second stage can be written as
0 0.5 1 1.5 2
Dilution rate (h–1) F • X1 F • X2 dX2
ⳮ l2 • X2 ⳱
Ⳮ 123
V2 V2 dt
Figure 5. Comparison of steady-state behavior of a chemostat 123 123 CELL
123 (51)
(solid lines) and a chemostat with recycle (dotted lines) using the CELLS IN CELLS OUT GROWTH ACCUMULATION
following parameters: YX /S ⳱ 0.5 g DCW/g substrate, lmax ⳱ 1.0
hⳮ1, KS ⳱ 0.02 g/L, S0 ⳱ 30 g/L, C ⳱ 2, and ␣ ⳱ 0.5. and
368 BIOREACTORS, CONTINUOUS STIRRED-TANK REACTORS

F' S0' and product formation could be studied in the second


F F
S0 CSTR.
S1
X1 Adding an additional feedstream to the second stage
(Fig. 6) provides the opportunity to introduce more of the
limiting nutrient, other nutrients required for growth or
product formation, inducers, or inhibitors. The material
balances on cell mass and substrate on the second stage
F2 can be written as equations 53 and 54, respectively, with
Volume V1 Volume V2 X2 steady-state solutions given in Table 5. In equation 53, F⬘
S2 is the volumetric flow of the second feedstream, which is
D1 = F/V1 assumed to be sterile (X⬘ ⳱ 0).

Figure 6. Two-stage chemostat system with possibility of a sepa- F1


rate feed to second reactor. X F1 Ⳮ F⬘ dX2
V2 1 ⳮ X2 Ⳮ l2 • X2 ⳱
123 V2
14243 123 dt
CELLS IN 123 (53)
CELLS CELL
FROM GROWTH ACCUMULATION
1 STAGE 1
OUT

0.8
The dilution rate for the second stage is given by D2 ⳱ (F1
τk = 1 Ⳮ F⬘)/V2, and the concentration of the limiting nutrient in
Conversion, X

0.6 τk = 0.5 the second feed is S⬘0.

0.4 F1 F⬘
S S⬘ F1 Ⳮ F⬘
V2 1 Ⳮ V2 0 ⳮ S2
123 123 V2
14243
0.2 τk = 0.1 SUBSTRATE SUBSTRATE
SUBSTRATE
IN FROM IN FROM
OUT
STAGE 1 SECOND FEED
0 l2 • X2
0 1 2 3 4 5 ⳮ ⳱ dS2
n YX/S
123 dt
123 (54)
CONSUMPTION
Figure 7. Substrate conversion for first order reaction in n ACCUMULATION
FOR GROWTH
CSTRs in series.

Feeding additional substrate to the second stage allows for


F • S1 F • S2 l2 • X2 more growth to occur. In addition, the dilution rate in the
ⳮ ⳮ ⳱ dS2 second stage is larger than the maximum specific growth
V2 V2 YX /S
123 123 123 dt
123 rate of the organism because the second stage has a con-
SUBSTRATE SUBSTRATE CONSUMPTION
ACCUMULATION tinuous feed of cells.
IN OUT FOR GROWTH
(52)
Enzyme Reaction. Enzyme reactions may also be carried
with steady-state solutions shown in Table 5. A growth out in multiple CSTRs. The performance of CSTRs in se-
model on the limiting substrate must be adopted to further ries approaches that of a single PFR while maintaining the
complete the system description. Minimal cell growth will good mixing characteristics of the stirred-tank reactor (37).
occur in the second stage if no additional substrate is Considering a first-order enzymatic reaction for substrate
added, because the majority of substrate is consumed in conversion (v ⳱ k ⳯ S), the reactor design equation for a
the first stage. Thus, non-growth-related cellular behavior single CSTR can be written as

Table 5. Steady-State Solutions to Material Balances for a Two-Stage Chemostat


Cell mass Substrate
First stage l1 ⳱ D1 X1 ⳱ YX /S(S0 ⳮ S1)

冢 冣
X1 D2
Second stage l2 ⳱ D2 1 ⳮ X2 ⳱ YX /S(S1 ⳮ S2)
X2 l2
F1 • X1
冢 冣
YX /S F1 F⬘
Second stage with additional feedstream l2 ⳱ D2 ⳮ X2 ⳱ S1 Ⳮ S⬘ ⳮ D2S2
V2 • X2 l2 V2 V2
where D2 ⳱ D1 ⳱ F/V1 where D2 ⳱ (F1 Ⳮ F⬘)/V2
BIOREACTORS, CONTINUOUS STIRRED-TANK REACTORS 369

S0 ⳮ S kd Thermal denaturation constant, hⳮ1


s⳱ (55)
k•S kLa Overall mass transfer coefficient, hⳮ1
kP Specific product degradation constant,
Assuming that there is no volume change upon reaction, hⳮ1
the conversion in the single CSTR is given by
KS, KI, KP Model parameters for cell growth
s•k Km, KS, K⬘m, K⬘S Model parameters for enzyme kinetics
X⳱ (56) m Maintenance coefficient g/g DCW ⳯ h
1Ⳮs•k
Ndiscs Number of discs in a disk-stack
Then for a system of n CSTRs in series with equal volumes centrifuge
and reactor conditions (constant k) the conversion in the OD660 Optical density at 660 nm; measure of
nth CSTR is given by equation 57 and depicted in Fig- cell concentration
ure 7. P Product concentration, g/m3
QAGIT Agitation heat input, W
1
Xn ⳱ 1 ⳮ (57) QEVAP Heat loss by evaporation, W
(1 Ⳮ s • k)
n

QEXCH Heat removal by heat exchanger, W


QLOSS Heat loss to surroundings, W
NOMENCLATURE
QMET Metabolic heat generation, W
Abbreviations qO2 Specific oxygen uptake rate, g O2 /g
DCW * h
CER Carbon dioxide evolution rate
qP Specific product formation rate, g/g
CIP Clean-in-place DCW * h
CSTR Continuous stirred-tank reactor R Ideal gas constant, J/mol * K
OTR Oxygen transfer rate
r2, r1 Outer and inner radii for centrifuge, m
OUR Oxygen uptake rate
RBATCH Biomass productivity for batch
PFR Plug flow reactor fermentation, g DCW/m3 ⳯ h
RQ Respiratory quotient, RQ ⳱ CER/OUR RCSTR Biomass productivity for CSTR, g
WFI Water-for-injection DCW/m3 ⳯ h
S Substrate concentration, g/m3
Symbols
S0 Inlet substrate concentration, g/m3
DHRXN Heat of reaction, 0 endothermic, 0 T Bulk fluid temperature, K
exothermic, kJ/mol
TC Coolant temperature, K
DQSENS Net sensible heat input, W
tCYCLE Batch cycle time, h
(RCSTR)M Biomass productivity at XM and DM, g
DCW/m3 ⳯ h tM Mixing time, h
A Area, m2 tTURNAROUND Lumped batch turnaround time, h
C Concentration factor in cell recycle U Overall heat transfer coefficient, W/m2
system ⳯K
C* Equilibrium dissolved oxygen u Particle velocity, m/h
concentration, g/m3 ulim Limiting particle velocity for a
CL Dissolved oxygen concentration, g/m3 separator, m/h
Cp Heat capacity, J/kg ⳯ K V Fluid volume, m3
D Dilution rate, hⳮ1 v Reaction rate,
Da Dimensionless Damkohler number vmax Maximum reaction rate
DC Critical dilution rate, hⳮ1 wH20 Mass flow rate of water, kg/h
DM Dilution rate associated with maximum X Cell concentration, g DCW/m3
RCSTR at fixed S0, hⳮ1 X0 Inlet cell concentration, g DCW/m3
dP Particle diameter, m Xi Inoculum cell concentration, g DCW/m3
E Enzyme concentration XM Cell concentration associated with DM,
E0 Initial enzyme concentration g DCW/m3
Ea Activation energy, kJ/mol YHi Cell mass produced per heat evolved, g
F Volumetric flow rate, m3 /h DCW/kcal
g Gravitational acceleration, 9.8 m/s2 YP /S Product yield on substrate, g/g
k2 Catalytic rate constant, hⳮ1 YX /O2 Cell yield on oxygen, g DCW/g O2

Das könnte Ihnen auch gefallen