Chapter 17
Chapter 17
Chapter 17
Bioprocess Engineering
and Technology
Thus from a total of (10 + 12.5) = 22.5 l, at the output 12.5 l needs to be recycled. (Note that the overall material
balance is not affected by the recycle.)
100 l h1
50 g l1
5 g l1
50 g l−1 R
5 g l−1 1 2 50 g l
−1
10 g l−1
Knowing CA0, k and D, the value of CA can be determined. We will return to this system when studying the kinetics
of cell growth.
Example 3 Let us consider an unsteady state situation where the accumulation term is nonzero. We have the same
CSTR where the input concentration of A is increased at time t = 0 to CAf from CA0. The material balance is
dCA/dt = D(CAf − CA) − kCA
This is a differential equation which needs to be solved for the given boundary conditions. Here we solve for the
simple case where there is no reaction term. The above case is left as an exercise for the interested student. When there
is no reaction we have
dCA / dt = D(CAf − CA ) or dCA / (CAf − CA ) = D dt.
However for material balance purposes we need to consider the main elements which participate in a significant way
in the process. We will develop here a basic technique to solve the set of material balance equations, which if required
can be extended to more elements as well.
Consider a system of microbial cells at SS. This way of formulating the system allows us to set the accumulation term
to zero and reduce the material balance equations to simple algebraic equations. In this system substrate is consumed and
biomass produced and these become the input to and output from the system, respectively.
Figure 17-5 shows the various flows going in and out of the system which is at SS. As all flows are shown as coming
in, the output flows would turn out to have negative values.
There are a total of seven flows consisting of two substrate flows (representing the carbon and nitrogen sources),
biomass, product, oxygen, carbon dioxide and water. The chemical formula for the flows will be written on a C-mole
basis which helps in setting up the material balances. This means that the formula will have one carbon atom only and
the number of atoms of the other elements are adjusted accordingly. Thus glucose, which is a typical carbon source, has a
chemical formula of CH2O (and not C6H12O6) and biomass will be represented as CH1.8O0.5N0.2 (and not C10 H18 O5 N2)
which is a fairly good approximation, representative of the elemental composition of a range of microbial cells. For exam-
ple the difference between yeast cells and Escherichia coli in terms of the relative ratio of the different elements present
in these cells is less than 5%. However, more complex chemical formulas may be used, if required, to represent biomass.
In order to simplify the material balance further, elemental balances on carbon, hydrogen, oxygen and nitrogen will
be done. Since elements do not participate in the reactions taking place, we can set the reaction terms equal to zero.
Thus the material balance reduces to
Input = Output
The general formula for a flow Fi is designated as CHaiObiNci. Thus if F1 (i = 1) is the flow of glucose (having the formula
CH2O) then a1 = 2, b1 = 1 and c1 = 0 and if F2 is the flow of biomass having the formula CH1.8O0.5N0.2 then a2 = 1.8, b2 = 0.5
and c2 = 0.2. Combining the elemental balances with the overall material balance which is ∑Fi = 0 we get the following
four balances:
C balance: ∑Fi = 0, (17.2)
H balance: ∑aiFi = 0, (17.3)
O balance: ∑biFi = 0, (17.4)
N balance: ciFi = 0. (17.5)
Note that we have seven flows and hence seven unknowns and only four independent equations. Thus three flows
need to be determined experimentally for defining the complete system. In order to simplify the solving of the four
simultaneous equations we follow the standard practice of multiplying each equation by a factor and then adding the
equations to get a simple equation. This multiplication factor can then be appropriately chosen to reduce the number of
unknown flows in the final equation.
Thus multiplying the four balance equations [(17.2)–(17.5)] by λC′ , λH′ , λO and λN, respectively, and adding we get
ΣlC Fi + ΣlH a i Fi + ΣlO bi Fi + Σl N ci Fi = 0.
Input profile
Output profile
CA
F1 SYSTEM F4
(Substrate (Substrate
CA0 C-source) N-source)
t0 Time F2 F3
(Biomass) (Product)
Figure 17-4 Unsteady state profile of concen- Figure 17-5 System of microbial cells at steady state with various
tration of A in a CSTR. flows.
777
17.3 Material Balance•
Rearranging we get
ΣFi (lC + a i lH + bi lO + ci l N ) = 0.
The expression (λc + aiλH + biλO + ciλN) can be represented by γi. Thus the final material balance becomes
ΣFiγi = 0.
Now the values of λC′ , λH, λO and λN can be carefully chosen so that certain γi values become zero. Thus choosing
λH = 1 and λO = −2, we get
γH2O = 2 λH + λO = (2 − 2) = 0.
(Since water does not have any carbon so there is no λc term and also c7 = 0 because no nitrogen is present.) Similarly
choosing λC = 4 would set γ CO2 = (λC + 2λO) equal to zero.
If the nitrogen source is ammonia gas then choosing λN = −3 would set γN = 0 (γN = λN + 3λH). However, if the nitrogen
source is complex, for example glutamate (CH1.8O0.8N0.2), then to get λN equal to zero we can use the equation
g N = lC + 1.8lH + 0.8lO + 0.2l N = 0,
or
Fsg s + Fx g x + Fpg p + Fog o = 0,
where subscripts s, x, p and o represent carbon source, biomass, product and oxygen, respectively. [Note that the γ values
of F4´ F6 and F7 (nitrogen source, carbon dioxide and water) are zero and hence they do not appear in the final material
balance equation.]
Of these four unknowns only three need to be determined experimentally in order to determine the complete stoi-
chiometry of the reaction. Let us illustrate this by simple examples.
Example 4 Consider the case when no product is being formed so that Fp = 0. Thus the material balance equation
becomes Fsγs + Fxγx + Foγo = 0.
We now define yield coefficient Yx/s which represents the amount of biomass formed per unit amount of substrate
consumed. Yield coefficient is thus a ratio having no units and can be represented as
Yx/s = C-mole of biomass formed/C-mole of substrate consumed.
Similarly other yield coefficients can be defined, for example Yx/o which is the ratio of biomass formed to oxygen
consumed. (This procedure is similar to choosing the ‘basis’ of the material balance as 1 C-mole of biomass formed and
calculating all the other flows with respect to this basis.)
The material balance equation thus becomes (on dividing by Fx throughout)
Fsg s / Fx + g x + F0g 0 / Fx = 0,
or
g x / Yx / s + g x + g o / Yx / o = 0.
Suppose cells are grown on glucose and ammonia as carbon and nitrogen sources, respectively, and the yield coef-
ficient is measured experimentally as 0.5 g/g (i.e. grams of biomass formed per gram of glucose consumed). Then the
oxygen yield coefficient Yx/o can be calculated from the above equation. We have
g s = 4 + 2 − 2 = 4(formula CH 2 O);
g x = 4 + 1.8 + 0.5 × 2 − 0.2 × 3 = 4.2(formula CH1.8 O 0.5 N0.2 );
g O = −4.
778 • Chapter 17/Bioprocess Engineering and Technology
The molecular weights of biomass and glucose on a C-mole basis are (12 + 1.8 + 0.5 × 16 + 0.2 × 14) = 24.6 and
(12 + 2 + 16) = 30, respectively.
Thus the yield coefficient can be calculated on a C-mole/C-moles basis as Yx/s = (−0.5/24.6)/(1/30)
Yx/s (C-mole/C-mole) = (−)0.6 C-mole/C-mole,
where the negative sign indicates that biomass is formed while substrate is consumed. Substituting values in the above
equation we get
Yx/o = −1.7 C-mole/mole of O2.
A negative value signifies that while biomass is produced oxygen is consumed in the reaction. To determine the values
of the other flows in this reaction we can go back to the original balance equation. Thus a carbon balance gives
Fs + Fx + FCO2 = 0
Dividing by Fx we obtain
1 / Yx / s + 1 + 1 / Yx / CO2 = 0.
a positive value indicating that both biomass and carbon dioxide are produced in the reaction. These values can be used
to calculate the ratio of oxygen consumed to carbon dioxide produced which is called the respiratory quotient (RQ)
given by YCO2/O:
YCO2/O = Yx/O/Yx/CO2 = 1.7/1.5 = 1.13 (with a negative sign which is being ignored since RQ is a ratio).
Similarly a nitrogen balance gives
F2C2 + F4C4 = 0
or
Fx × 0.2 + FN = 0,
Yx/NH3 = −5 C-mole/C-mole.
Thus all the flows can be determined with respect to one C-mole of biomass formed.
Can you now determine whether water is produced or consumed in this reaction?
Example 5 Consider another simple case where there is anaerobic growth as happens in yeast fermentations producing
ethanol. Then since Fo = 0, we have
Fsγs + Fxγx + Fpγp = 0.
If the sources of carbon and nitrogen are glucose and ammonia, respectively, then on dividing through by Fx we
obtain.
4/ Yx/s + 4.2 + 6 Yp/x = 0,
where Yp/x is the product yield per unit biomass on a C-mole basis and γp = (4 + 3 – 0.5 × 2) = 6 (since the formula of
ethanol is CH3O0.5). The above equation allows us to determine the product yield if the biomass yield is known. For
example, if Yx/s = 0.1 C-mole/C-mole then substituting this in the above equation, we get
Yp/x = 5.97 C-mole/C-mole
It often makes sense to determine Yp/s, the product yield per unit substrate. Dividing the above equation by Fs, we get
g s + Yx / sg x + Yp / sg p = 0.
The maximum product yield Yp/s is achieved when there is no cell growth and the biomass yield is zero (e.g. while
running an immobilized column of yeast cells which converts glucose to ethanol). This is given by
g s + Yp / cg p = 0.
Yp/s = −2/3 C-mole/C-mole.
Thus even when no biomass is produced one-third of the carbon flux goes to produce carbon dioxide.
The student may have perceived by now that the above choice of γ has in effect given us an electron balance of this
complex reaction of substrate consumption with biomass and product formation. In aerobic growth the ultimate electron
acceptor is oxygen where each molecule of oxygen can accept four electrons, while in anaerobic growth the product
is the final electron acceptor. The student is also expected to try out the material balances using different carbon and
nitrogen sources. For example, if glutamate is used as the nitrogen source, the value of λN = −21 and thus γx = 4 + 1.8 −
0.5 × 2 − 0.2 × 21 = 0.6. Thus, even though the equation remains
Fsg s + Fxg x + Fpg p + Fog o = 0,
If NH3 is the nitrogen source (λN = −3), γx = 0 + 1.8 + 0.5x − 2 + 0.2x−3 = 0.2 γ0 = −4 and γCO = 4.
2
For a case of no product formation we have
Fx (0.2) + Fo (−4) + FCO2 (−4) = 0.
Dividing by Fo we get
0.2 Yx/o − 4 − 4RQ = 0,
which allows us to relate the (RQ) with the biomass yield of oxygen.
Example 6 Consider a reactor where air is bubbled at the rate of 0.5 vvm (vvm represents air flow rate per unit volume of
culture per minute, thus 0.5 vvm mean 0.5 l of air per min per l of culture). If the inlet oxygen and carbon dioxide concen-
trations are 21 and 0% and outlet concentrations are 19.5 and 1.7%, respectively, determine, the rate of biomass formation.
Solution: Basis 1 l of culture
Since 22.4 l of gas at NTP contains 1 mole of the gas (ignoring temperature and pressure corrections). Oxygen
consumed per = 0.5(0.21 − 0.195)/22.4 = 3.35 × 10−4 mol.
CO2 produced per min = 0.017 × 0.5/22.4 = 3.795 × 10−4 mol.
RQ = 1.13.
Substituting in the previous equation:
0.2 Yx / o − 4 − 4 RQ = 0 (since RQ is -ve)
Yx / o = (−4 × 1.13 + 4) / 0.2 = −2.65 C-mole/mol −1 .
For spontaneous reactions the free energy change (ΔG) is always negative, thus the reaction term in this equation
would always be negative. However the system free energy denoted by the accumulation term may change either way
(or remain constant) depending on the input and output values.
For enthalpy balance we can write
ΔH reactants − ΔHproducts = qmet
where qmet is the sensible heat produced during the reaction. During growth, substrate is consumed and biomass (and
products) formed and the difference in their enthalpies gets reflected in the metabolic heat qmet generated during the
process. We can safely approximate the enthalpy content of oxygen, carbon dioxide and water to zero and calculate ΔH
for substrate, biomass and product by the standard heats of formation.
In this regard an interesting observation which has been confirmed empirically for a large number of substrates is
that if we choose λC = 4, λH = 1,λO = −2 and λN = 0 then the γ values obtained are proportional to the standard heats of
formation, that is
∆H i ∝ g i′,
Use of γ instead of γ′ allows use of previous material balance equations. Thus FsΔHs is the enthalpy content of the
substrate of which ηH(FSΔHS) is the enthalpy content of the product (in this case biomass) and (1−ηH)FSΔHS is the
metabolic heat generated in the process. The student can easily verify that
FsΔHs (1 − ηH) = 460Fo
We can now formulate two boundary conditions: first that ηH<1 which states that the enthalpy content of the
p roducts can never be higher than the reactants (enthalpy limitation) and second a carbon limitation, that is Yx/s < 1
which states that the carbon content of the product can never be higher than the reactants (if the nitrogen source
contains no carbon).
If we now plot these two boundary conditions by looking at how the yield coefficient (Yx/s) changes with increasing
values of γs we get Yx/s = ηHγs/γxThus as γs increases (for ηH<1), the upper bound of (Yx/s) increases till (Yx/s) reaches 1 after
which it enters the carbon limitation phase (Figure 17-6).
The figure simply demonstrates that when γs values are low, the substrate enthalpy content is also low. Thus since
the enthalpy content of the product has to be lower than reactants (enthalpy limited condition), the upper limit of Yx/s
increases linearly with increasing γ till it reaches 1 after which it does not increase further (carbon-limitation).
This switch from enthalpy-limited to carbon-limited would take place around a γs value of 4.2 (since γx = 4.2).
Let us now see how the plots of nH and qmet change with γs (Figure 17-7).
Figure 17-7 shows that the enthalpy efficiency declines as γs increases beyond 4.2 because once the upper limit of
Yx/s is reached the excess energy per C-mole of the substrate is simply lost as metabolic heat. This is obvious from the
equations
hH = Yx / sg x / g s and qmet = 115Fsg s (1 − hH ).
Experimental results from a range of fermentations show that the typical values of both nH and Yx/s never exceed 0.6,
much less than the upper limit of 1 set by the boundary conditions. As a matter of fact most fermentations which are
enthalpy-limited have a nH value between 0.5 and 0.6 and those which are carbon-limited a Yx/s value between 0.5 and
0.6. These values give us a very useful handle to predict within a fairly narrow range the expected performance of any
bioreactor with respect to its biomass and product yield, oxygen requirements, metabolic heat generation, etc., all of
which are useful parameters to know while designing a bioprocess.
We end this section by noting that, it is the free energy change (ΔG) and not ΔH which drives a reaction forward.
Thus ideally we should have dealt with free energy changes. However correlations for ΔG have much more scatter (and
hence error) and it is best to consult tables of the standard free energies of formation for accurately determining the
change in free energies. A typical correlation with around 10–15% error is
∆Gi = 94.4 × g i′ + 85.5.
Enthalpy Carbon
limitation limitation
H
qmet
Yx/s
4.2 s
s
H qmet
The student may use this correlation to determine the free energy change in a fermentation. In anaerobic fermentations
no oxygen is consumed, hence the enthalpy change (as predicted by the earlier correlation) is zero. This is indeed true
to a large extent because the metabolic heat generation is fairly low in an anaerobic system. (This is the reason why
anaerobic waste treatment plants do not require any heat removal devices and hence are easy to install and run).
However the free energy change is nonzero and it is the large entropy increase which drives the reaction forward.
Can you use the correlation given to determine the relationship between change in free energy and carbon dioxide
production in an anaerobic system?
17.4 Bioreactors
In any fermentation process the bioreactor plays a central role in determining the process efficiency. Even with recom-
binant products where stringent quality control implies that downstream processing is the major cost component, it is
the bioreactor performance which determines product yields.
Any vessel with facilities for aeration and agitation can be used as a bioreactor. Additionally it must meet the require-
ments of aseptic operation and provide the cells with a controlled environment conducive to growth and product for-
mation. Traditionally the vessel of choice has been the stirred-tank bioreactor which consists of a vessel with a vertical
rotating shaft with agitator blades (Figure 17-8).
Since the vessel must be sterilized, it must meet the requirements of pressure vessel design (since steam is used under
pressure to sterilize the bioreactor). Thus the material of construction is usually 4–5 mm thick stainless steel, which is
also resistant to the acids that are typically produced during fermentation. The height to diameter ratio (H/D) varies
from 1 (for small reactors) to 3 (for larger vessels). A high H/D ratio provides for a smaller footprint (space saving),
ease of construction and better mixing since the agitator blades need to be proportionally larger as the vessel diameter
increases. However tall tubular reactors with a high H/D ratio often lead to oxygen starvation in the gas phase which
affects oxygen transfer rates. The other factor which is critical for large reactors is heat transfer since the metabolic heat
generated during growth has to be removed. Cooling coils are used to carry cold water or a mixture of glycerol and water
if the temperature of the coolant has to be less than 0 °C.
The design of the agitator blade plays a crucial role in oxygen transfer and mixing. Traditionally the flat blade
impeller, the so-called Rushton stirrer (Figure 17-9) was used. This stirrer provides good radial mixing and does not get
‘flooded’ with air bubbles even with high air-flow rates. However it provides poor bulk mixing and this becomes a prob-
lem for fungal fermentations. This is because fungal broths shows non-Newtonian behavior where the viscosity changes
with the shear force. Thus the air bubbles get channeled through the center of the fermentor where the viscosity is
low (due to high mixing) effectively starving the media which are
Stirrer close to the walls of oxygen. To prevent this, agitators with better
shaft seal
bulk mixing characteristics have been designed, such as the Scaba
Aseptic
inoculation pipe
agitator and the Prochem Maxflow agitator (Figure 17-9). When
cells are shear sensitive the agitator is designed like a marine pro-
Working level
peller which gives good axial mixing.
Often the problems of heat and oxygen transfer cannot be
Impeller addressed with the conventional stirred-tank design. Airlift reac-
tors are then used to provide better oxygen transfer. These reactors
are modifications of bubble column reactors which have been used
Baffle in beer production. In bubble columns the mixing is provided by
the stream of bubbles entering the reactor at the bottom which
helps to reduce construction and operating costs. In airlift reactors
the liquid also rises with the bubbles through the ‘riser’, disengages
with the gas phase and comes down through the downcomer. This
Sampling ‘downcomer’ may be internal or external to the reactor. The cir-
P culation loop set up helps in improving heat and mass transfer
point
rates. Other specialized bioreactors are required for specific needs.
Thus photo-bioreactors need to have a large specific surface area
Sterile air line
(i.e. surface area per unit volume) since the growth of cells is
Air sparger dependent on the incident sunlight. Thus instead of a large tank,
Drain the cells are grown in tube banks (i.e. tubes arranged parallel to
point each other) and the material of construction is Plexiglas which is
Figure 17-8 Schematic of a stirred-tank bioreactor. transparent to sunlight. The media (usually seawater) enters with
783
17.5 Kinetics of Microbial Growth•
(A)
(B) (C)
Figure 17-9 Design of various agitators: (A) Scaba agitator, (B) Rushtor stirrer, (C) Prochem Maxflow agitator.
a small inoculum, flows through the tubes like in a plug glow reactor and provides a high cell density at the outlet. Often
a small fraction of the outlet cells is recycled to provide a continuous inoculum. Many novel bioreactors have been
designed, like the pulsed column bioreactor which combines a pneumatic or mechanical pulsing of the reactor medium
with a bubble column design. This helps in bubble break up thereby increasing the surface area and oxygen transfer rates.
Since the shear forces are low, this setup can be used for highly aerobic but shear sensitive organisms. However, most
of the designs remain at the lab bench as they have not been taken up by industry which still prefers conventional and
time-tested designs for large-scale operation.
Many interesting kinetic equations have been proposed to define this function, from the very simple to extremely
complex. However given the typical statistical errors associated with kinetic measurements especially in biological sys-
tems (things are changing a bit now!), there is little advantage gained in choosing complex kinetic equations to describe
cell growth. Thus unless the experimental data is sufficiently precise so as to distinguish between simple and complex
models of growth kinetics, the choice is fairly clear. One of the most popular (and simple) models which assumes that
the specific growth rate is a function only of the substrate concentration (usually assumed to be the carbon-source) is
the Monod equation
µ = µms/(Ks + s),
where µm is the maximum specific growth rate of the cells and Ks is a measure of the affinity of the cells towards the
substrate. The equation is similar to the Michelis Menten equation for enzyme kinetics and thus has the shape as shown
in Figure 17-10.
This similarity with enzyme kinetics is because cell growth is limited by the substrate uptake rate which in turn is
governed by the enzymes responsible for transporting the substrate inside the cells. However note that while Vmax in
Michelis Menten kinetics can be increased by increasing the enzyme concentration, µm is an intrinsic property of the
cell denoting the maximum rate of growth when it is not limited by substrate availability.
Example 7 Consider the situation when cells are growing at µm. [This assumption only requires that the substrate
concentration s is much larger than Ks (from Figure 17-10 you can see that µ ≈ µm when s >> Ks).]
If µm = 0.5 h−1determine the time required for cells to increase 10-fold in a bioreactor?
We have dx/dt = μmx or
xf t =t
∫ xi
dx / x = ∫
t=0
m m dt ,
ln(x f / x i ) = m m t,
where xi is the initial cell concentration at time t = 0 and xf is the final cell concentration at time ‘t’. The above equation
is referred to as the logarithmic form of the growth equation.
We can also write the above equation in the form
xf /xi = emt
which is referred to as the exponential form of the growth equation. Given that xf /xi = 10, we have
ln 10 = µt,
t = (ln 10)/0.5 = 4.6 h.
We can use the above equation to derive the relationship between the time required for the biomass to double, the
so-called ‘doubling time’ (td) and specific growth rate:
ln(xf /xi) = µt.
B A
m Given that in time td the value of xf = 2 xi, we get
ln 2 = µtd,
td = (ln 2)/µ.
Note that these two parameters are inversely correlated; the higher
the specific growth rate, the smaller is the doubling time.
The concept of doubling time is practically useful only in c ircumstances
when cell numbers can be counted and when they are undergoing binary
fission. Even then the cell number of a population increases continuously
and not in discrete jumps as is the picture obtained when one visualizes
one cell dividing into two and then into four and so on. This is because
there is a large population of cells that is dividing asynchronously in a bio-
KS
reactor. Such a number-based analysis also becomes tedious when we need
s
to know the cell number after time intervals that are not a whole multiple
Figure 17-10 Monod kinetics relating specific of the doubling time (say after one and a half doublings). Thus ‘doubling
growth rate (µ) to substrate concentration (s). time’ method of calculations finds little use in growth kinetics other than
785
17.5 Kinetics of Microbial Growth•
Yx/s = Δx/Δs.
Since the reactor volume is a constant, we can write
Yx/s(s0 − s)=x − x0, (17.6)
where Yx/s is the yield coefficient which is the stoichiometric ratio between biomass production and substrate consumption,
s0 is the initial substrate concentration, x0 the initial biomass concentration and s and x are substrate and biomass con-
centration at any given time. The increase in biomass concentration with time can be found using the biomass balance:
Accumulation = Input − Output ± Reaction.
Since the input and output terms are zero we have accumulation = reaction or
dx/dt = µx. (17.7)
Note that this is not a definition but a material balance on biomass which assumes no volume changes and no inputs
or outputs into the reactor.
The third equation is the kinetic equation governing the relationship between specific growth rate and substrate
concentration
µ = µms/(Ks + s). (17.8)
This is a simplification since in a reactor with cell growth the specific growth rate will be affected by more than one
parameter (e.g. changes in pH by product accumulation reduction in the concentration of other nutrients in the culture
medium, etc.)
Solving the three equations above, we have
s = s0−1/Yx/s (x − x0).
Therefore
dx / dt = { m m [ s0 − 1 / Yx / s (x − x 0 )]x } / { K s + [ s0 − 1 / Yx / s (x − x 0 )]}.
or
∫A dx / x + ∫B dx / (s Y 0 x/s − x + x 0 ) = m m ∫ dt,
where
A = (s0 Yx / s + x 0 + K s Yx / s ) / (s0 Yx / s + x 0 ),
B = K s Yx / s / (s0 Yx / s + x 0 ).
Recycled cells
Input Output
Figure 17-13 Schematic of a plug flow reactor. (substrate) (cells)
788 • Chapter 17/Bioprocess Engineering and Technology
Accumulation = Input − Output ± Reaction,
S Biomass balance V(dx/dt) = F × 0 − Fx + µxV,
X where F is the flow rate of both input and output and x
is the biomass concentration in the reactor, which is also
equal to the biomass concentration in the output (since the
reactor is well mixed). At steady state, if we denote the
biomass concentration by x , we get
Fx = m x V or m = D [ because D = F / V ]. (17.11)
Using the above equations we can predict the biomass and substrate profiles at steady state at different dilution rates:
D = m = m m s / (K s + s ) or s = DK s / (m m − D)
and similarly
x = Yx / s [ s0 − DK s / (m m − D)].
Figure 17-15 show a plot of s and x against D. Note that the value of x does not change significantly when the values
of D are low. This is because of the typically low values of Ks which lead to low values of s . Thus for a wide range of D
values, the substrate gets almost completely consumed, leading to a constant biomass concentration in the reactor. (The
student may check out as to how this pattern changes when the Ks values change.)
As the dilution rate is increased, µ rises till it reaches the upper limit given by µm. At this point the rate of production
of biomass cannot keep pace with the output flow and the cells start getting washed out of the reactor. This rate of ‘wash
out’ can be determined by the following equation:
dx/dt = mmx − Dx. (17.13)
which is a negative term given that D > µm. Slow washout is often used in lab-scale reactors to determine the value of µm
by running the CSTR at values of D greater than µm and plotting ln x vs. time. This gives a slope of (μm−D) from the
equation ln x = (μm−D)t.
The CSTR can also be used to determine the values of YX/S, µm and Ks by measuring outlet biomass and substrate
concentrations for different values of D.
789
17.5 Kinetics of Microbial Growth•
A plot of 1/D vs. 1/ s (double reciprocal plot) has a slope of Ks/µm and an intercept of 1/µm. Yx/s can be determined
from a plot of x vs. (s0 − s).
/ s = 1 / Yx / s + m / m .
1 / Yxobs true
(17.14)
/ s vs. 1/µ would thus give a straight line with a slope ‘m’ and intercept 1 / Yx / s .
A plot of 1 / Yxobs true
Such a graph can easily be plotted using a CSTR where µ = D and Yx / s = x /(s0 − s).
obs
[Note that if we introduce the concept of m (and hence a varying Yx/s) in a batch reactor, solving the integral in order
to get x vs. time is no longer possible algebraically and numerical techniques will have to be used.]
The plot of x vs. D would also change if the maintenance requirements are significant (Figure 17-16) while the plot
of s vs. D would remain unchanged.
The concept of maintenance is also useful in describing stationary phase cultures where substrate is consumed but no
biomass is produced.
In order to link up with the stoichiometric relationships which were developed in the previous section we need to
understand that the substrate consumed for maintenance purposes is essentially converted to carbon dioxide and thus
oxygen is required for this process. We can thus write analogously for oxygen consumption,
/ o = 1 / Yx / o + m o / m ,
1 / Yxobs true
(17.15)
because the balance equations are correct for both situations. Substracting
Eq. (17.17) from Eq. (17.16) and substituting in Eqs. (17.14) and (17.15)
we get
gs (ms/µ) = 4(mo/m) D
Example 9: Calculate the product yield (YP/S) for ethanol using an anaerobic culture of yeast growing on glucose as a
carbon source at two specific growth rates, µ = 0.5 h−1 and 0.05 h−1 given Yxtrue
/ s = 0.4 C-mole/C-mole and ms = 0.2 h .
−1
Solution: Since
/ s = 1 / Yx / s + m s / m ,
1 / Yxobs true
However one model which combines the two models mentioned earlier has found applicability to a wide range of
products. The model is defined by the equation
qp = αm + β.
First proposed by Luedking and Piret in 1959, it has been used by many authors to explain the kinetics of product
formation of lactic acid and many other primary metabolites which are not completely growth dependent.
The model has a growth-associated and a non-growth-associated term both of which can be determined using a
CSTR. We have a product balance:
V dp/dt = (αm + β)x − Dp.
At steady state
(am + b)x = Dp or p / x = a + b / D.
A plot of p / x vs. 1/D would give the values of α and β from the intercept and slope, respectively. [The student may
try and determine the product profiles in a batch reactor for different values of α and β. Thus when β = 0 and the product
is completely growth-associated, the product profile would be similar to the biomass profile (since YP/X is a constant).]
Solving the above three equations we get m = Dx / (x + xw ) which tells that µ < D in this system. Substituting this
in the kinetic equation, we obtain
DYX / S (s0 − s ) / [ YX / S (s0 − s ) + x w ] = m m s / (K S + s ).
This is a quadratic equation in s , which can be solved to get
Increasing s and hence x .
xw
x Example 10: Given the steady-state wall growth is 1 g l−1.
Determine outlet biomass concentration (x ) at a dilution rate
of 0.5 h−1, given YX/S = 0.4 g/g, µm = 0.5 h−1 and Ks = 0.1 gl−1 and
s0 = 10 gl−1
Solution: Substituting the values in the above equation, we have
0.5 × 0.4 (10 − s ) 0.5 s x
= .
0.4 (10 − s ) + 1 0.1 + s
D
⇒ s = 0.38gl −1 ,
Figure 17-18 Plot of biomass vs. dilution rate for CSTR
with wall growth. which gives x = 3.8 gl −1.
Note that if there was no wall growth we would have observed
xout ‘wash out’ since D = μm (in this case). But wash out never takes
place since the cells adhering to the wall keep growing and pro-
ducing cells. Figure 17-18 shows the plot of x vs. D for different
levels of wall growth.
(The reader may try out how maintenance requirements
would change the above figure.)
which is similar to the earlier formulation for internal recycle, with A = (1 + R − Rg); this value is less than 1 since g ≥ 1.
(The student may check that a biomass balance around the CSTR also gives the same result.)
We note that all the above methods, viz. CSTR with wall growth, CSTR with internal or external recycle, have a
mathematically similar analysis because in all cases the biomass concentration in the outflow is lower than the biomass
concentration in the reactor thus leading to a situation where μ < D. These reactors can thus run at higher dilution rates
without having a wash-out situation. However, lower µ values lead to low values of YX/S (if we were to consider mainte-
nance requirements) and also enhanced oxygen requirements per unit biomass formed.
Example 11: Waste treatment plants often prefer low concentrations of biomass at the outlet since they cause problems
of disposal. Consider a CSTR with external recycle requiring the outlet biomass concentration to be ≤ 10 g 1−1. Given
that s0 = 50 g l −1 , YX / S = 0.4 g/g , m s = 0.3 h −1, D = 0.2 h−1, g(concentration factor) = 10, what is the recycle ratio at which
the reactor should be run?
Solution: Assuming that Ks is very small, xout = YX / S (s0 ) since s values are around Ks and thus {(s0 − s ) ≈ s0 }
/ s = 10 / 50 = 0.2 g/ g.
Yxobs
Now
/ s = 1 / Yx / s + m / m .
1 / Yxobs true
Thus 44% of the cells need to be recycled in order to achieve the desired concentration of cells in the outflow. Note
that if Ks values were large then µm needs to be known and the solution to this problem becomes quite complex involving
a trial and error method.
FSF The reactor slowly fills up and the process is stopped when the vessel
is filled up or the desired product/biomass concentration is achieved. The
reactor may be partially emptied so that the residual culture serves as a
starting inoculum and the feed started once again in the case of ‘repeated
fed-batch’ culture. The slow feed has many advantages, chief among
them being the ability to control the residual substrate concentration in
the reactor and hence the specific growth rate of the cells. Excess sub-
strate concentration often leads to poor product formation (the ‘crabtree’
effect). To avoid this, the initial substrate concentrations are kept low
in the batch phase. However, if only this substrate were available for
Figure 17-22 Schematic of fed-batch culture. growth this would lead to low final biomass concentrations. Feeding sub-
strate slowly into the reactor overcomes this problem thereby helping in
the achievement of high biomass densities. Also the rate of feed can be
adjusted so as to get the desired specific growth rate which is best suited for product formation.
where F is the feed rate, SF is the feed concentration and X = xV. If the specific growth rate is low then the residual
substrate concentrations are also low (given low Ks values).
Thus all the substrate added is completely consumed by the growing cells giving
FSF = mxV / Yx / s + mX.
Consider a situation when the feed rate is a constant and maintenance requirements are negligible. We get
mx = FSF Yx / s / V,
which is a constant. This is a situation of linear growth where ‘µx’ is a constant and hence µ declines as x increases
throughout the cultivation period.
If we consider maintenance effects to be significant then with increasing biomass a larger proportion of substrate
would be utilized for maintenance and the growth rate would decline slowly till it reaches zero and all the substrate is
utilized for maintenance. In such a case
FSF = mXmax.
Note that this is also the situation when the biomass levels are the highest. If µ is to be held constant, the feed rate
needs to be increased exponentially so as to match the desired specific growth rate.
where D = F/V.
If the substrate concentration of the feed were to be the same as the initial batch concentration (and given that the
residual substrate concentration in the reactor is very low), we can define a pseudo-steady-state condition where dx/dt = 0
or m = D. However unlike a CSTR the dilution rate is not a constant since D = F / V = F / (Vo + Ft) for a constant feed
795
17.6 Heat Transfer•
rate. D thus continuously declines leading to a decline in µ. We can consider this situation analogous to the linear growth
described previously, where instead of a linear increase in biomass concentration the volume increases linearly. Note that
in both cases since µ declines the residual substrate concentration declines as well. However, assuming low values of Ks, this
decline does not add significantly to the biomass produced which is given by (since s0 s )
x = YX / S (s0 − s) ≈ Ys0 .
Fed-batch techniques are extremely useful while doing high cell density cultivation (HCDC) and producing second-
ary metabolites, where substrate starvation is an important trigger for product formation.
Example 12: Determine total product formed in a fed-batch culture with constant feed given F = 100 l h −1 , SF = 200 gl −1 ; V = 10, 000 l, Yx / s =
−1 −1
h , SF = 200 gl ; V = 10, 000 l, Yx / s = 0.4 g/g, s0 (initial substrate concentration in the batch phase) = 20 g1−1. The production formation
kinetics is given by qp = 0 if ; m ≥ 0.1h−1; qp = β = 0.05 when 0.02 ≤ m ≤ 0.1 h−1; qp = 0 when m ≤ 0.02 h−1. (Assume Ks to be
very small.)
Solution: Let the feed begin only after cells have grown and the initial substrate is almost exhausted, that is
x = Ys0 = 8 gl −1 .
Thus product formation would start with the start of feed and continue till µ falls to 0.02 h−1. This fall in µ is because
of increase in x. Using the substrate balance again and substituting µ = 0.02 h−1, we get x = 40 g l−1.
At this point the value of µ falls below 0.02 h−1 and no further product will be formed. The rate of product formation
is given by
q p = 1 / x(dp / dt) = b
p t
dp / dt = bx or ∫ dp = ∫ bxdt
p0 0
where po is the product concentration at the end of the batch phase (which can be considered to be zero. The rate of
increase of x is linear and is given by
dx / dt = mx = FSF YX / S / V = 0.8gl −1 h −1x = x 0 + 0.8t.
Substituting for x in the integration given above and putting appropriate boundary conditions, we get
32
which is the amount of product formed. The problem would become complex if maintenance requirements are introduce
in the material balance, because Yx/s would no longer be a constant.
R = ∆X1 / K 1 + ∆X2 / K 2 ,
(T1 − T2 ) = (q / A)(∆X1 / K 1 + ∆X2 / K 2 );
T2
which is similar to V = iR.
The numerical values can be substituted to determine q/A.
Similarly, since voltage drop is proportional to resistance, we
Figure 17-24 Temperature gradient for steady-state have
heat flow through two slabs.
T1 − Tinterface /(T1 − T2 ) = (∆X1 / K 1) / (∆X1 / K 1 + ∆X2 / K 2 ),
q = ko A(Tb − Two ) / d o ,
q = k i A (Twi − Tc )d i ,
where ko and ki are the conductivities of these films (of liquid) and δo and δi are the thicknesses. Since it is impossible to
experimentally measure this film thickness we can couple this with the conductivity of the film and define a film heat
transfer coefficient h given by
ho = ko /d o and h i = k i / d i
there by getting
q = ho A(Tb − Two ),
q = h i A(Twi − Tc ),
q = k A(Two − Twi ) / ∆X,
where k is the conductivity of the wall (typically stainless steel which is the material of construction of the cooling tube)
and ΔX is the thickness of the cooling tube. Adding up the resistances we have
1 / U = 1 / ho + ∆X / k + h i ,
where U is defined as the overall conductivity and h0 and hi are the external and internal ‘film heat transfer coefficients,’
respectively. We can now write the heat balance as
q = U A ∆T = U A(Tb − Tc ),
where q is the heat removed by the cooling coil, A is the surface area of the coil and ΔT is the temperature difference
between the fermentor culture medium and the cooling liquid. The student can try and derive the above expression
using basic algebraic techniques which involve using the three equations given earlier to get rid of the two unknowns,
namely, Two and Twi.
However this temperature difference (Tb − Tc = ΔT) is not a constant since the temperature of the cooling liquid rises
as it flows inside the fermentor (here it picks up the heat generated by the growing culture). This change in the cooling
coil temperature (ΔTc) changes ΔT between the coil and fermentor medium (Figure 17-26). Thus an average value of
ΔT needs to be used while using the above equation.
This average ΔT can be determined if we know the ΔT at the inlet and the outlet of the cooling coil. We have
q = m Cp ∆Tc ,
where m is the mass flow rate of the cooling liquid and Cp is its heat capacity.
If we consider the heat exchange in the small shaded portion in the Figure 17-26, we have
dq = U dA ΔT,
where dA is the area of the shaded part. Also from the temperature increase of the cooling liquid we have
dq = mCp d(ΔT)
since the change in ΔT is because of the rise in temperature of
the cooling liquid. Thus
mCp d(ΔT) = U dA ΔT.
Rearranging and integrating across the whole length of the Tb
coil, we get Temp
Tc
mCp ∫ d(∆T)/ ∆T = U ∫ dA
T
therefore
q = mCp (ΔT1 − ΔT2).
Substituting for mCp in the above equation we get
q = U A[(ΔT1 − ΔT2)/ ln ΔT1/ΔT2].
The term within brackets is referred to as the log mean temperature difference (LMTD) and the use of this logarith-
mic mean instead of the arithmetic mean of the temperature differences (ΔTavg) gives a more accurate estimate of the
heat transfer rate. Thus
q = UALMTD.
Example 13: Estimate surface area of cooling required for a reactor growing yeast cells at 30 °C, given that
m = 0.2 h −1 , x = 20 g l −1 ; V = 1000 l; hH = 0.6; U = 100 W/m −2 K −1, cooling coil liquid is water which enters at 15 °C and
leaves at 20 °C. Also estimate the flow rate of cooling water in the coil?
Solution: Rate of biomass generation is (0.2 × 20 × 1000)/24.6 = 162.6 C-mole/h.
Now
qmet = 162.6 ∆H X (1 − n H ) / hH = 6 × 104 kJ h −1 .
The total heat generated (qT) can be assumed to be 1.25 times this value (if 80% of the heat generated is due to
metabolic heat):
qT = 1.25 × 6 × 104 kJ h −1 = 20.83 kJs −1 .
LMTD = (∆T1 − ∆T2 ) / ln ∆T1 / ∆T2 ,
where ΔT1 = (30 − 15) = 15 °C and ΔT2 = (30 − 20) = 10 °C (since the reactor temperature is 30 °C and water enters at 15 °C
and leaves at 20 °C).
LMTD = (15 − 10)/ ln(15/10) = 12.33 °C.
(Note that this value is slightly lower than the arithmetic mean of temperature difference 12.5 °C)
We have seen that
q = UALMTD
⇒ 20.83 × 10 (Js ) = 100 W m −2 K × A × 12.33
3 −1
⇒ A = 16.9 m 2 .
⇒ m = 3571 kg h −1 .
A more accurate analysis would require independently estimating the other heat flows in and out of the fermentor.
These functional relationships need to be determined empirically. However, since there is a physical basis to these
relationships we postulate that the form of these relationships require that they be dimensionally homogenous. This
means that the LHS and RHS of this equation have the same units in terms of the four fundamental units, viz. mass
(M), length (L), time (t) and temperature (T). The requirement of dimensional homogeneity imposes constraints on the
nature of the equation that can be formulated and thus reduces the number of constants that need empirical determina-
tion. The strategy outlined below is called the Buckingham Pi method and though we are applying it now to determine
the relationship between h and other variables, it has applicability in a wide range of other cases. We define a power law
relationship
h = Ck a D b m c Cdp v e r f ,
where C is a dimensionless constant and a, b, c, … are the exponents. Note that if there were no constraints on this e quation
we would need to empirically determine the value of the seven constants. The dimensional units of the v ariables are
h = [ Mt −3T −1 ]; k = [ MLt −3 T −1 ]; D = [ L ];
m = [ ML−1t −1 ]; Cp = [ L2t −2 T −1 ]; v = [ Lt −1 ];
r = [ ML−3 ].
The equation in terms of the dimensional units is (C being a constant has no units)
[ Mt −3 T −1 ] = [ MLt −3 T −1 ]a [ L ]b [ ML−1t −1 ]c [ L2t −2 T −1 ]d [ Lt −1 ]e [ ML−3 ]f .
for L : 0 = a + b − c + 2d + e − 3 f ; (17.19)
for T : − 1 = − a − d. (17.21)
We have four equation but six unknowns and thus we can express four of the above exponents in terms of the other
two. Retaining the variables ‘d’ and ‘f’ (this makes the job easier), we get
from Eq. (17.21), a = (1 − d);
Nu = C (Pr)α(Re)β,
where the constant C and the values of α and β need to be determined empirically. We have thus reduced the number
of constants that need empirical determination from seven to three. Typical values of C, α and β are available in liter-
ature helping us to correlate the above dimensional numbers. Thus the Dittus Boelter equation for liquid flowing inside
pipes is:
(Nu) = 0.0265 (Re)0.8 (Pr)0.3 .
where NA is the flux of component A per unit area, D is the diffusivity and (dCA/dx) is the concentration gradient. This
is similar in form to the equation governing heat transfer. We are concerned with mass transfer primarily because cells
growing inside a fermentor need oxygen which has to be supplied by bubbling air/oxygen into the fermentor.
For the transfer of oxygen to the cells, oxygen has to first diffuse through the bubbles into the fermentor medium
and then into the cells from the liquid medium. The equilibrium relationship governing gas solubilities in a liquid is
governed by Henry’s Law:
pA = HcA
where pA is the partial pressure in the gas phase, cA is the solubility (referred to as the dissolved oxygen concentration)
in the liquid phase and H is Henry’s constant. The value of H is a function of pH, temperature, ionic strength and pres-
ence of dissolved solutes and hence varies from medium to medium. When air is bubbled inside liquid (either water or
fermentor medium) so that equilibrium is reached, the dissolved oxygen concentration reaches its saturation value c*
given by p A = Hc A ∗ which is typically in the range of 1.0 ± 0.2 mmoles 1−1.
Let us now consider the situation where oxygen is transferred from bubbles to medium containing growing cells and
simultaneously taken up from the medium by these cells for respiration. A material balance on the liquid phase for
oxygen gives
V (dcL/dt) = OTR − OUR,
where V(dcL/dt) represents the rate of accumulation of oxygen in the liquid,
OTR is the oxygen transfer rate from the bubbles to the liquid and OUR is the
Gas Liquid interface
oxygen uptake rate by the growing cells given by
OUR = µx/YX/O.
Pb
At steady state dcL/dt = 0 and we have OTR = OUR
Note that the dissolved oxygen concentration at steady state given by cL is
lower than c* (the saturation value), thus allowing continuous oxygen transfer
Pi
from the gas to the liquid phase.
Ci To determine OTR we consider the resistance to diffusive flow of oxygen due
to two films, one on the gas side causing the partial pressure of oxygen (which is
CL the driving force) to drop from pb (the bulk gas phase oxygen partial pressure) to
pi (the partial pressure at the interface) (Figure 17-27).
Similarly the film on the liquid side causes the dissolved oxygen concentration
Figure 17-27 Oxygen gradient due to to fall from ci and cL, where ci and cL are the dissolved oxygen concentrations at
diffusion of oxygen from bubbles to cul- the interface and bulk liquid, respectively. (This formulation is similar to the for-
ture medium containing growing cells. mulation for heat transfer.) We can now write the equations for mass transfer as
801
17.8 Mass Transfer•
N = K G A(pb − pi ),
where N is the flux of oxygen, KG is the gas side film mass transfer coefficient, A is the surface area of the bubbles. Also
N = K L A(ci − c L ),
The specific surface area (a) can be estimated if we know the volume of gas held up in the reactor (Vg) and the radius
of the bubbles (r). We therefore have
ε = Vg/VT,
where ε is called the gas hold up ratio. Also the specific surface area of a bubble is given by its area per unit volume, that
is
a = (4pr 2 ) / (4pr 3 / 3) = 3 / r.
Thus the specific surface area per unit reactor volume is given by Vg × a / VT = 3e / r .
[Since we have a bubble size distribution we need to use ravg, a mean radius of the bubbles where the concept of ‘Sauter
mean radius’, rsm, is used (rsm = Σri 3 / Σri 2 ). The student can check that the Sauter mean radius gives an accurate measure
of the specific surface area.]
Correlations are available for KL in terms of the diffusivity D and other variables in the form of dimensionless num-
bers. Similarly correlations for a are also available in literature. However, for practical purposes since both KL and a affect
the OTR, we use correlations which estimate KLa as a single parameter. A typical correlation is
K L a = C (P / V)a (u s )b ,
where C is a constant, P is the power consumed due to agitation, V is the reactor volume and vs is the superficial gas
velocity (given by the gas flow rate divided by the cross-sectional area of the reactor).
The power consumption for agitation can be related in turn to other variables using dimensionless numbers. Thus
P = f (N, D, ρ, g, µ), Laminar Transient Turbulent
where N is the rpm of the agitator, D the diameter of the agitator, ρ
the liquid density, g is gravity and µ is the viscosity of the liquid. The Po
dimensionless numbers relating the above variables are (the students
can derive this from the Buckingham Pi method described earlier)
(P / N 3 D 5 r) = c(rND 2 / m)a (N 2 D / g)b .
The LHS of the equation is the Power number (P0), the term
(ρND2/μ) is the modified Reynolds number for agitated liquids (Re) Re
and (N2D/g) is the Froude number.
Typically ‘baffles’ are used in reactors to break up vortex f ormation. Figure 17-28 Plot of Power number vs. Reynolds
These are thin strips of metal stuck to the inside wall of the reactor number.
802 • Chapter 17/Bioprocess Engineering and Technology
(in the perpendicular direction), which break up the circular motion of the liquid and thus cause turbulence. Under
these circumstances the Froude number has little effect on the power consumption. A plot of Power number vs. Reynolds
number is given in Figure 17-28, where three zones can be identified during agitation. These are:
1. Laminar zone: It is the zone where po declines with increasing Re.
2. Transient zone: It is the zone where a complex relationship exist between po and Re.
3. Turbulent zone: It is the zone where po is a constant (and hence independent of Re).
Most fermentors are run in the turbulent zone and therefore the Power number is a constant which is independent of
the Reynolds number. We thus have
P = Po (N 3 D 5 r).
This relationship can also be derived from basic principles if we note that the power consumption is proportional to
the kinetic energy (KE) it imparts to the liquid while mixing. Thus the mass of liquid an agitator blade displaces while
rotating is proportional to ‘ρND3’ (where D3 is proportional to the volume displaced by the agitator blades, so ND3 is the
volume displaced per unit time) and the velocity it imparts to this liquid is ‘ND’ (which is the tip velocity of the agitator
blade). Thus the KE imparted is
1
2
mv2 ∝ rND 3 × (ND)2 = rN 3 D 5 .
if the aeration rate is kept constant per unit volume of reactor (i.e. 1 vvm), then the superficial gas velocity is given by
v = vvm × D3/D2 = D × vvm.
Thus v2/v1=10. This increase is very high and could lead to problems of foaming and flooding of the agitator. Thus the
aeration rate in vvm is typically reduced to 0.5 vvm as the scale increases. Let us assume the following correlation for KLa:
K L a ∝ (P / V)0.7 (v s )0.3 .
Also let the system be running in the turbulent zone and hence
Po = constant and P ∝ r N 3 D 5 .
Thus
K L a ∝ (r N 3 D 5 / D 3 )0.7 (vvmXD)0.3 = N 2.1 D1.7 vvm 0.3 .
803
17.8 Mass Transfer•
we can write
dCL / dt = K L a(C* − CL ) − K L a(C* − CL )
⇒ dCL / dt = K L a(CL − CL ).
∫ dC L / (CL − CL ) = ∫ K L a dt. CL
E
Controller System Output
Measuring
device
Figure 17-30 Schematic of feedback control.
805
17.9 Measurement and Control of Bioprocess Parameters•
the set point value. This over correction (which is called the
‘overshoot’) will then require further corrective action and a Final set
situation may arise where the cycles of overshoots continue point
indefinitely and the pH oscillate continuously around the set C
point. On the other hand, a very slow response of the con-
troller may lead to an unacceptable delay in the correction of B
the error. The possible responses to a step change in the set
A
point are given in Figure 17-31. Initial set
It should be noted that a slight ‘overshoot’ while correct- point
ing the error is usually acceptable and forms a basic feature of
the response curve. However, the oscillations should lessen
in amplitude quickly enough (a quarter rate decay is usu- Time
ally acceptable, i.e. the amplitude of the oscillations should
decline by a quarter in each cycle). Figure 17-31 Response of a system with feedback control
to a step change in the set point.
17.9.2 Controller Characteristics
As stated earlier the controller response is a function of the error (ε). Typically the response is given by
The first term Kcε represents the proportional component of the response, that is the response is proportional to the
magnitude of the error. When only this term is present the controller is called a proportional controller. It should be
noted that this proportional response cannot increase indefinitely with increasing ε since it would soon reach its upper
limit (of 100% response). For higher values of ε the response thus saturates at its upper limit. Therefore the response is
proportional only for a range of ε values. This range is called the proportional band (PB). We can relate Kε to PB by the
following equation.
Kc = 100%/PB.
Thus a higher value of Kc (which is also referred to as the controller gain) implies a narrow proportional band.
The problems associated with having only a proportional control is that the controller response declines as the ε
value falls thereby leading to ‘offset’ error which is the permanent deviation from set point. To understand this consider
a simple case of a growing culture producing acids which lower the pH. The proportional control would correct for this
fall in pH by adding base at a rate proportional to the deviation from the set point. However, when this deviation is low,
the rate of base addition would fall till it matches the acid production rate leading to a permanent offset error in the set
point pH. A solution to this lies in increasing the proportional ‘gain’ which can, however, lead to an oscillatory response.
To counter this, integral response is introduced where the response is dependent on the integrated sum of errors over
time. This removes both offset error and reduces the possibility of oscillatory response. The third component of the
response is the derivative response which allows a faster response time. However random noise or any error having sharp
fluctuations can cause an inappropriate response and hence derivative control needs to be used with care.
Taken together they form the proportional-integral-derivative (PID) response which is a standard feature of most
controllers today. The values of Kc, τI and τD need to be finetuned so as to get the best response and these are dependent
on the characteristics of the system being controlled. Setting these values is referred to as controller tuning.
These values give the standard quarter decay response; however, lower values may be chosen if a more conservative
response is required.
806 • Chapter 17/Bioprocess Engineering and Technology
Slope = S
Step
Variable input Output
t=0 TD
t Figure 17-32 Process reaction curve of a system.
Process reaction curve: In this method the response of the system to a small step input ‘m’ is measured in the absence
of any control. A typical response is shown in Figure 17-32.
A tangent is drawn at the point of inflexion of the response curve. The point where this tangent cuts the x-axis is
used to determine the value of TD, the delay time in the response. The normalized slope is given by S, where S is the
slope of the tangent/m.
The values of TD and S are used to determine the controller characteristics where
K c = 1.2 / TD S; t I = 2 TD ; t D = 0.5 TD .
The process reaction curve may also be used to model the system response by considering it equivalent to a delay time
response plus a first order response (see Box 17-1)in which case the system response to various controller settings can be
predicted for different kinds of errors in a quantitative fashion. From these an appropriate choice of controller settings
can easily be made.
Major advances have taken in recent years in process control techniques. Thus ‘adaptive’ control is now used where
the controller settings are changed as the process changes. For example, as the culture grows from low to high cell den-
sity the system characteristics with respect to fluctuations in the dissolved oxygen concentration change. Hence the
controller settings are tuned in real time.
Feed-forward control is another method often used to counter the effect of large delay times. If good models of the
process are available then the possible effect of input errors can be estimated and corrective action initiated even before
the system output shows the error. Controller training can also be done using complex computer techniques like artificial
neural networks or support vector machines.
C = C0 (1 − e − Dt ),
where the initial concentration at time t = 0 is zero and the step input is C0. Such a response is called a first-order response.
If two CSTRs are connected in series, the output would be a second-order response. The student can check that the typical
process reaction curve given in the earlier section can be simulated by the combination of a time delay response followed
by a first-order response.
807
17.10 Sterilization•
17.10 Sterilization
Sterlization is the process of getting rid of contaminating agents like bacteria, viruses, etc. from the system so that we
can use it for growing the desired microorganisms. Thus the bioreactor vessel and medium need to be sterilized. Also
all inputs to the system like air, media feed, acid/base need to be sterile and aseptic operating conditions need to be
maintained. The inoculum needs to be free of contaminating agents aswell.
Air sterilization is a fairly straightforward process nowadays because of the availability of filters which retain all the
contaminating agents with 100% efficiency. This is possible because polymers can now be designed (usually PTFE-
based) where the pore size can be controlled very accurately. Similarly the reactor vessel can be sterilized with steam at
high pressure (121 °C for 20 min). Media sterilization, however, poses problems because the constituents undergo degra-
dation during the sterilization process and hence the trade off between nutrient quality and degree of sterilization needs
to be well-understood and optimized.
∫ dN / N = ∫ −kdt.
⇒ ln(N 0 / N f ) = −kt.
The term ln N0/Nf is called the ‘del’ factor denoted by the symbol ‘∇.’ The ∇ value represents the required
degree of sterilization. For example, if the initial cell count is 103 cells ml−1 and the media volume is 1000 l then
N0 = 103 × 103 × 103 = 109 cells. Nf is chosen typically as 0.001 which implies that the probability of 1 cell surviving
post-sterilization is 1/1000 (or that 1 fermentor in a 1000 would not be completely sterile because 1 cell would survive
the sterilization process). Then ∇ = ln 109 / 10 −3 = ln 1012 = 27.6 . To achieve this desired ∇ value a time–temperature
regimen is required, which is given by
∇ = Ae−ΔE/RTt.
If the temperature is a constant, the time can be easily determined (since a constant temperature implies a constant
value of k). However, media which is usually sterilized in holding tanks (or the fermentor vessel itself) usually has a
heating phase (where the temperature rises), a holding phase (where the temperature is kept constant) and a cooling
phase (where the temperature is brought down to normal). Typically since the fermentor or the holding vessel cannot
withstand high pressures (> latm gage) the holding temperature is 121 °C. The overall ∇ is thus
∇total = ∇heating+ ∇holding + ∇cooling.
If the temperature profiles for heating and cooling phase are available, the above equation can be integrated to obtain
the heating and cooling values. Numerical integration can also be done by calculating the ‘k’ values for short time
intervals where the temperature can be assumed to be a constant and then adding the values. Thus
Σ∇ = ΣkΔt.
A simplified calculation protocol can be used if one assumes linear heating and cooling rates and also that significant
contributions to the ∇ value are obtained only above 100 °C.
Table 17-1 gives the ‘k’ values for B. stearothermophilus spores for different temperatures above 100 °C. Assuming a
1 °C/min−1 rise/fall in temperature, the ∇ values for different temperatures are also given. Thus at T = 115 °C, the ∇
value has been calculated by assuming that the temperature rose from 100 °C to 115 °C in 15 min.
808 • Chapter 17/Bioprocess Engineering and Technology
Table 17-1 The ‘k’ values for B. Stearothermophilus spores for different temperatures
T(°C) k (min−1) ∇heating/cooling
100 0.019 —
101 0.025 0.044
102 0.032 0.076
103 0.040 0.116
104 0.051 0.168
105 0.065 0.233
106 0.083 0.316
107 0.105 0.420
108 0.133 0.553
109 0.168 0.720
110 0.212 0.932
111 0.267 1.199
112 0.336 1.535
113 0.423 1.957
114 0.531 2.488
115 0.666 3.154
116 0.835 3.989
117 1.045 5.034
118 1.307 6.341
119 1.633 7.973
120 2.037 10.010
121 2.538 12.549
122 3.160 15.708
123 3.929 19.638
124 4.881 24.518
125 6.056 30.574
126 7.506 38.080
127 9.293 47.373
128 11.494 58.867
129 14.200 73.067
130 17.524 90.591
An example of how to calculate sterilization time by the simple method is given below.
Example 15: The initial cell count presterilization is 104 cells ml−1 and the media volume is 10,000 l. The heating from
100 °C to 121 °C (the holding temperature) takes 15 min and the cooling down to 100 °C takes 20 min. Calculate the
holding time.
Solution: We have
∇total = ln 104 × 104 × 10 3 / 10 −3 = ln 1014 = 32.2.
If the heating was at a rate of 1 °C min−1, then from 100°C to 121 °C, ∇heating = 12.549 (Table 17-1).
However, since the heating took only 15 min we have
∇heating = 12.549 × 15/21 = 8.96.
Similarly
∇cooling = 12.549 × 20 / 21 = 11.95.
We know that
∇total = 32.2 = ∇heating + ∇holding + ∇cooling.
Therefore
∇holding = 11.29.
Since k at 121 °C is equal to 2.538 (Table 17-1) the holding time for adequate sterilization is given by ∇holding = kt
or
809
17.11 Media Design•
t = ∇holding/k = 4.44min.
Note that the heating and cooling phase contribute significantly to the overall sterilization process. Thus, in order to
avoid oversterlization (and hence excess nutrient degradation) we need to include the effects of heating and cooling as
well. Continuous sterilizers are used where the culture medium flows through a heat exchangers which rapidly increases
its temperature to high values for a short period. This high-temperature–short-time-period regimen reduces nutrient
degradability. To understand this we need to look at the basic sterilization equation:
∇ = Ae − ∆E / RT t
⇒ ln ∇ = ln A − ∆E / RT + ln t.
If we consider a constant degree of sterilization (∇ = constant), a plot of ln t vs. I/T gives a straight line of slope ΔE/R
(Figure 17-33).
Points on this line represent different time–temperature regimens which give the same degree of sterility. Lines drawn
parallel and above this line would represent larger values of ∇.
Now consider nutrient degradation (especially of thermolabile components in the media) which is also a first-order
relationship
ln (x0/xf) = kt,
where x0 is the concentration of these nutrients in the initial medium and xf is the final concentration after sterilization.
The first-order rate constant ‘k’ also has an Arrhenius dependence on temperature but the ΔE values in this case are
much lower (20–30 kcal g−1 mole−1). Thus a plot of ln t vs. I/T for constant degree of nutrient degradation would have
a lesser slope than that for constant ∇ (Figure 17-15). Lines parallel and below represent lesser degree of nutrient deg-
radation. From the figure we can deduce that choosing short time intervals and high temperature allows to reduce the
degradation and hence improve nutrient quality while keeping ∇ values constant.
not always possible with complex media since each ingredient contains a range of elements. An approximate analysis of
commonly used media ingredients like molasses (which is used as a carbon source) or corn steep liquor (used as a nitro-
gen source) is available in literature. Since the carbon source usually serves the dual purpose of supplying carbon as well
as energy to the cell, the material balance needs to be formulated with care.
Typical carbon source are: cerelose (commercial glucose), glycerol, cane or beet molasses, oils (soyabean, corn and
cottonseed), corn starch, dextrins, whey (65% lactose) and alcohols (e.g. methanol).
Typical nitrogen sources are: Corn steep liquor, soyabean meal or flour, dried distillers solubles, yeast extract, fish
meal, peanut meal, etc.
Other than carbon and nitrogen, cells need minerals, trace elements and vitamins for growth. Additionally in
fermentors we need to add buffers (for pH control) precursors, inhibitors or inducers to regulate the metabolism of the
cells and antifoams to reduce foaming. The design of the final media thus becomes an exercise of juggling with many
variables each of which may act independently or in concert to effect the fermentor performance. The traditional
method of optimizing media, where one variable is varied and the others kept constant, is not very useful in these
circumstances. These have been replaced by statistical techniques which allows the simultaneous changing of many
parameters (usually the concentration of media ingredients) in order to get the best media design.
The testing of all possible combinations of media concentrations is called the complete factorial design. Since this
would lead to an unacceptably large number of experiments, only certain critical combinations are tested in what is
called a fractional factorial design. Two such fractional factorial designs are described below.
The results of the experiments for which A was low (L) are averaged together and subtracted from the average of the
results of the experiment when A was high (H) to get the effect of A:
Effect of A = (ΣH A − ΣL A ) / 6.
If the statistical variation were to be negligible, the effect of the dummy variables should be zero. Therefore, the
average effect of the dummy variables is taken as a measure of the statistical error. An ‘F’ test which compares the square
of the effect of the concerned variable with that of the dummy variables gives us a relative index which defines the
importance of each variable. The variables with the highest F values can now be defined as the key variables and chosen
for further study.
Therefore, for commercial purposes, as well as for further study and analysis, a series of separation techniques are exe-
cuted in succession to purify a product from the myriad of possible impurities. A fermentation broth typically consists of
a complex aqueous mixture of cells, soluble extracellular products, soluble/insoluble intracellular products and uncon-
verted substrates to name a few. When solutions carrying particulate substances, having densities greater than the
solutions, are allowed to stand for long, the particles settle down under the action of gravity with a velocity which is a
complex function of its size and density.
There are many different types of separation methods used for biological materials, based on their molecular p roperties,
differences in size, charge, relative solubility, affinity for a particular chemical, etc. This section introduces the basic fea-
tures of downstream processing, the methods of bioseparation following an industrial bioprocess. The strategy used for
any particular case depends on a lot of factors like (i) broth characteristics (viscosity, product concentration, impurities,
etc.); (ii) final product purity and concentration and (iii) the desired form of the product (crystallized product, dried
powder, concentrated liquid, etc.)
where KL is a constant and ϕ is the volume fraction of particles. KL depends on particle shape, thus for spherical particles
KL = 2.5 and for rod-shaped particles KL varies from 20–1200.
For more concentrated solutions, Eilers derived the following relationship
hrelative = [1 + K L Φ × 0.5 (1 − Φ / Φ max )−1 ]2 ,
where Φmax is the maximum packing capacity (0.6–0.7) for the type of cells involved.
Example 16: Determine the shear stress for a suspension of spherical cells in water with a cell volume of 0.2 and a max-
imum packing density of 0.6 when the shear rate is 60 s−1.
814 • Chapter 17/Bioprocess Engineering and Technology
Solution: Since the cell volume is 0.2, we use Eilers equation to determine relative viscosity (ηrelative), in which, KL = 2.5
as the cells are spherical. With Φ = 0.2 and Φmax = 0.6, we get ηrelative = 1.89. Since the viscosity of water is 0.001 (Nm−2s)
at 20 °C we have ηsuspended = 0.00189. Substituting in the earlier equation, at dv/dx = 60s−1, we get t = 0.1134 Nm −2 .
where K is the consistency index (Nm −2 s n ) and n is the power law index. Both these constants can be measured by a
rheometer typically a Brookfields viscometer which involves measuring the torque required to rotate two concentric
cylinders at varying rates when the annulus is filled with the liquid whose viscosity needs to be measured. When η < 1
the liquid is referred to as a pseudo-plastic while dilatant fluids have η > 1.
Many fermentation broths follow the Bingham plastic model given by
t = t 0 + h(dv / dx),
17.13.3 Sedimentation
Sedimentation is the settling of substances in a simple gravitational field. (The settling achieved is enhanced by
centrifugal force, and in these cases faster rates of settling are obtained.) At a larger-scale sedimentation can be done
when cells have a tendency to adhere closely (coagulate) or to form multicelled aggregates (sometimes aided by pre-
treatment). Many flocculent yeast strains, for example, are in use in breweries and single-cell-protein production. In cell
stic
pla
am
Bingh
tic
p las
do
eu
Ps ian
ton
N ew
t
la tan
Di
Figure 17-35 Herschel–Buckley model
relating shear stress to shear rate for
dv/dx Newtonian and non-Newtonian liquids.
815
17.13 Downstream Processing•
suspension settlers, the rate of setting declines with increasing solid concentration. The interface between the cell-free
and the cell-bearing liquid moves downward according to the equation:
U s = kc − m ,
where Us is the setting velocity, c is the cell concentration, k is a constant and m takes a value between 1.7 and 2.6. This
is called hindered settling. (The principles governing free settling are given in the next section.)
Sedimentation finds use in sludge waste treatment for removal of water. Solid wastes are converted to cell-sludge by
biological treatment. This suspension can be treated by sedimentation to separate the free water and get concentrated
sludge which can be used as fertilizer.
17.13.4 Centrifugation
A centrifuge is a piece of equipment that accelerates the rate of sedimentation by rapidly spinning the samples, thus
creating a centrifugal force many times that of gravity. The force F, acting on a particle of volume Vp and density rp
located at a distance r from a point about which it is revolving with angular velocity ω is the difference between the
centrifugal force, mω2r, and the buoyant force (Vp ρω2r) exerted by the solution. That is F = mw 2 r − Vp rw 2 r, where Vp is
the particle volume and ρ the density of the solution. Since m = VPρp and the drag force can be expressed by Stokes Law:
Drag force = 6phRVr .
where η is the visocity of the fluid medium and Vr is the terminal particle velocity in the r direction and R is the radius
of the particle.
Integrating the above equation allows us to determine the time (T) required for a particle to move from a distance r1
from the center of rotation to r2. Thus
T = 9 / 2h / w 2 R 2 (rp − r) ln r2 / r1 .
Different particles will therefore take different times to settle through the same distance. A particle may be
c haracterized by its sedimentation rate. Every particle can have a sedimentation coefficient, s = V/ω2r, where V is the
terminal settling velocity and ω2r is the centrifugal force. This coefficient depends on the shape, size and density of the
particle and also on the viscosity of the liquid. This is the basis for differential centrifugation, separating different sized
particles using different centrifugal forces and different times.
Centrifuges can be of many types (serving different purposes) depending on their internal features (solids bowl, solids
effecting, nozzle, scroll, etc.) Preparative ultracentrifuges can be used in two distinct ways:
1. Zonal centrifugation in which particles are separated according to their sedimentation coefficients.
2. Equilibrium density gradient centrifugation where particles are separated according to their densities.
17.13.5 Filtration
This is the most common method of separating solid particles in suspension (like whole cells) from liquid media.
Separation is accomplished by forcing the liquid through a porous membrane where the solid particles are trapped. These
particles build up as a layer on the surface of this membrane called a ‘cake’ while the clear liquid goes through. To attain
high throughput rates the pressure drop for flow may be increased or the resistance to flow may be decreased. Most indus-
trial equipment decreases the resistance by increasing the filtration area. This is because increasing the pressure of liquid
tends to compact the ‘filter cake’ especially if it is made up of compressible material like cells. This in turn increases the
cake resistance and hence reduces the flow. In batch filters as more liquid flows through, the solids deposited on the filter
membrane increase thereby increasing the cake thickness. This in turn reduces the flow due to increased resistance of
the filter cake and the filtration has to be stopped after some time so that this cake can be removed from the filter. The
length of the batch cycle depends on the liquid flow rate, the pressure used, the amount of solids in the feed, etc. Filter
aids are often used to prevent blocking of the filter membrane by the solids, and also help in the formation of a porous
filter cake, which lowers the cake resistance and increases the length of the batch cycle.
We define Vs = 1 / A (dV / dt) When Vs is the flow velocity, A is the surface area of filtration and dV/dt is the flow
rate. We have
∆ PahVs L
where ΔP is the pressure drop, η is the viscosity and L is the cake thickness. The proportionality constant would depend
upon the porosity and particle size of the cake. The thickness of the cake (L) can be related to the volume of filtrate by
a material balance
LA(1−€)ρs = wV,
where € is the porosity, ρs is the density of solids in the cake, w is weight fraction of solids in the feed and V is the volume
of filtrate. Substituting we get
∆P = Kη(1/A dV/dt) wV/[A(1−€)ρs].
Rearranging and putting α = K/(1 − €)ρs, we get
1 / A dV / dt = ∆P / (ha w V / A),
where α is called the specific cake resistance.
Often the resistance of the filter medium (the membrane) needs to be incorporated into the calculations. For conve-
nience, this resistance is expressed in terms of an equivalent volume of filtrate:
(1 / A)dV / dt = ∆P / [h a w (V + Ve ) / A ].
Here Ve is the volume of filtrate necessary to build up a fictitious filter cake the resistance of which is equal to that of
the membrane.
Rearranging we get
(V + Ve )dV = ∆PA 2 dt / h a w.
which allows us to calculate the time required to collect a certain volume of filtrate. Typically, however, we need to
calculate α and Ve from experimental runs. For this we write
dt / dV = h a w (V + Ve )/ A 2 ∆P.
This gives a straight line plot between (dt/dV) and V if ΔP is kept constant. From the slope and intercept the values
of α and Ve can be calculated, which in turn can be used to predict filtration performance.
For small volume fermentations, a plate and frame filter can be used, which can be opened and the filter cake removed
after each batch operator. For larger volumes, continuous filters are required.
Various kinds of continuous filters are used, one of the most common being a rotary vacuum filter, where strings are
used to lift off the rotating filter cake which accumulates on the filter surface. The clear liquid passes through the surface
to the inside due to the vacuum applied.
17.13.6 Precipitation
Precipitation is usually done to separate and concentrate intracellular proteins after cell lysis. The two major methods
used are: (a) salting-out by adding inorganic salts such as ammonium sulphate at high ionic strength; (b) solubility
reduction at low temperatures by adding organic solvents such as acetone.
In the first process the ions added interact strongly with water, causing protein molecules to precipitate. The solubility
of the solution is given by
ln S / S0 = K s I,
where S is the solubility of the protein in solution, S0 is the solubility at I = 0, I is the ionic strength of solution and Ks is
constant depending on pH and temperature.
Alternatively, addition of organic solvents at low temperatures can precipitate proteins by reducing the dielectric
constant of the solution. The solubility of a protein as a function of the dielectric constant of a solution is given by
817
17.13 Downstream Processing•
ln S / S0 = K ′ (1 / D 2s ),
If we calculate the flows on a solute-free basis and also assume that the L and V phases are essentially immiscible then
we can write
L0 = L1 and V1 = V2 .
Thus the subscripts of L and V can be dropped. This also requires that the solute concentrations (x and y) be reported
as grams per liter of pure solvent (and not grams per liter of solution) since the mass flow rates of the L and V phases
would change as the solute is transferred from one phase to the other.
The component balance gives
Lx0 + Vy2 = Lx1 + Vy1.
The equilibrium relationship between the solute concentrations leaving the equilibrium stage can be written as
V1Y1 V2Y2
(1)
L 0X 0 L1X1
Figure 17-36 Schematic of single equilibrium
operation.
Lx0 X1 X2 Xn
Lxn1
1 2 n n1
Vyn2
Figure 17-37 Schematic of multistage equilib- Y1 Y2 Y3 Yn+1
rium operation with counter current flow.
818 • Chapter 17/Bioprocess Engineering and Technology
y1 = f (x1).
17.13.8 Chromatography
Chromatography is a process of separation of mixtures into its components by passing the fluid mixture through a bed of
adsorbant material. In the commonly practiced elution-type chromatography, a column is packed with adsorbed gel (or
porous solid/liquid phase immobilized on a solid). A fluid mixture is injected, followed by an eluent. Different solvents
interact differently with the adsorbent material. Solutes interacting weakly with the matrix pass out rapidly while those
interacting strongly exit slowly. The differential migration rates are used to separate different components. Some of the
important chromatographic methods are:
1. Adsorption chromatography (ADC) is based on the different adsorption of solute particles onto solid particles like
alumina and silica gel by van der Waal’s and steric interactions.
2. Liquid–liquid partition chromatography (LLC) is based on the different partition coefficients of solute molecules between
an adsorbed liquid phase and passing solution.
3. Ion-exchange chromatography (IEC) is based on the adsorption of ions on ion exchange resins by electrostatic forces.
4. Gel filtration chromatography (GFC) is based on the penetration of solute molecules into small pores of packing m aterial
on the basis of molecules shape and size.
5. Affinity chromatography (AFC) is based on specific chemical interactions between the solute molecules and bound
ligands.
6. Hydrophobic interaction chromatography (HIC) is based on hydrophobic interactions between proteins (or other solute
molecules) and the functional groups (e.g. alkyl residues) on the support matrix.
7. High pressure liquid chromatography (HPLC) is based on general chromatography except that it is done under high
liquid pressure which provides fast and high resolution.
17.13.9.1 Dialysis
Dialysis is a membrane-based process that separates low molecular weight solutes from a solution. The dialysis membrane
is selective and has a cut off value allowing solutes with molecular weight lower than the cut off to move through the
membrane from a high to a low concentration region. At equilibrium the chemical potentials of a diffusing solute on
both sides of the membrane are equal, that is
ma = m b
⇒ RT ln Ca Y a = RT ln Cb Y b
⇒ Ca Y a = Cb Y b ,
where Cα, Cβ are concentrations of the solute in the two phases and Yα, Yβ are the activity coefficients. For ideal
solutions
Ca = Cb .
Hence by varying the volume of one of the phases, the concentration of the solute in the other phase can be controlled.
where Σp i denotes the osmotic pressure due to ith component, Ci is concentration of the ith component, T is the
temperature and R is the gas constant.
In ideal cases B2i = B3i = 0, and the equation reduces to
p = Σp i = ΣCi RT.
The magnitude of pressure varies according to the concentrations of the solutes. As an example, a pressure of
30–40 atm is required for a 0.6 M salt solution. A major problem with RO membranes is due to increased concentration
of solutes close to the membrane surface, reducing solvent flow. This phenomenon is called concentration polarization.
where De is the effective diffusivity of the solute in liquid film (cm2), J is the volumetric filtration flux of the liquid (cm3/
cm2s) and C is the solute concentration (mol cm −3 ) .
This equation can be solved if the specific boundary conditions are known.
chemical and other purposes. The usual strategy to bring about crystallization is to direct a macromolecular solution
toward a supersaturated state by modifying the properties of the solvent or the solute. This can be achieved by as simple
a physical process as the change of temperature or pH. Alternatively, a change of salt concentration or the addition of a
precipitating agent can also bring about crystallization. Most processes use a combination of various factors like altering
pH and salt concentrations and addition of precipitating polymers. Out of the many approaches attempted so far, the
most successful has been the manipulation of salt concentration. The reason behind this is that, as the salt ions compete
with each other for hydrogen bonding with the solvent molecules, the increase in salt concentration forces the protein
molecules to self-associate to satisfy their electrostatic requirements. Polymers that dehydrate macromolecules (e.g.
polyethylene glycol) and organic solvents like ethanol and acetone also help crystallization in similar ways. However,
optimization of crystallization conditions is a matter of empirical experimentation.
Techniques used for drying also depend on the physical properties of the product, properties of the solid–liquid
system, the drying environment and heat-transfer parameters. The following processes are usually applied for drying.
1. A vacuum-tray drier consisting of heated shelves is mainly used for pharmaceutical products.
2. Freeze drying (lyophilization), a common method used for antibiotics, enzymes and bacterial suspensions, uses
sublimation as a method of removal of water from a frozen solution.
3. Rotary drum driers remove water by evaporation from a thin film of solution on the steam-heated surface of a rotating
drum and the dried product is scraped from the drum with a knife.
4. Spray dryers, used for heat sensitive materials, use an atomization technique where the product solution is sprayed
into a heating chamber through a nozzle where the solvent evaporates and the dry particles are separated using
cyclones.
5. Pneumatic conveyor dryers use a hot air stream to suspend and transport particles. They are also well suited for
heat-sensitive and easily oxidized materials.
4. Oxidation pounds are shallow reactors where mixed cultures grow. They require large land areas and can have
environmental side-effects.
Magnetic bead
for agitation
8% (w/v) alginate
solution containing cells
Drops
Gelation
(15 min)
2% w/v CaCl2
solution
H2O wash
30 min Drying 20 h
30 min
Biocatalyst
beads
Glucose
L
So
Glycolysis
Pos
Pyruvate
decarboxylase
Pyruvate
Mg2+, Thiamine pyrophosphate
Soi
Acetaldehyde + CO2
Alcohol dehydrogenase
NADH2
Ethanol
Poi
Sos 17.15.2 Production of Organic Solvents
17.15.2.1 Acetone/Butanol
Chain Weizman performed fundamental research on the
Support fermentation of Clostridium acetobutilicum for the production of
surface
Po acetone, butanol and ethanol. During World War I, acetone was
the product of interest, used for the production of the e xplosive
tri-
nitrotoluene. After the war, butanol became important
Biofilm
primarily for the production of nitrocellulose lacquers. Butyric acid,
Liquid film butanol, acetone and isopropanol are all obtained through clos-
Figure 17-40 Schematic representation of a biofilm tridial fermentation of starch, molasses, sucrose, wood hydrolysates
showing substrate and product profiles in the liquid film and pentoses. The relative proportions of each of these products
and biofilm. S0: subtrate concentration in bulk medium; depends on the strain used and the fermentation conditions.
S0i: substrate concentration in the biofilm liquid inter- Of the many fermentation products, only the acetone-butanol
face; S0s: substrate concentration on the support surface. part is of current economic interest. During the bioprocess, one
has to remember that butanol at less than 0.5% has no influence
on the cells, while at higher concentrations it causes damage to
the cell membrane and at concentrations above 1.3% b utanol production ceases. Autolysins are also produced some-
times which causes the cells to lyse.
the biomass is separated by filtration. Precipitation is usually accomplished by addition of Calcium hydroxide (lime) to
the fermentation broth. The precipitate is washed and treated with dilute sulphuric acid yielding an aqueous solution of
citric acid and CaSO4. After bleaching and crystallization, either anhydrous or monohydrate citric acid is obtained. The
demand for citric acid was 400,000 tons in 1999 ($ 1,400 million), which is expected to increase 3% annually.
17.15.5.2 Streptomycin
Aminoglycoside antibiotics (streptomycin, kanamycin, gentamycin) are produced in fermentors of 150,000 l capacity,
with optimal aeration, a temperature of 28–30 °C and a pH in the neutral range. The process is of 4–7 days duration.
826 • Chapter 17/Bioprocess Engineering and Technology
Glucose with starch is used as carbon source. Soya meal acts as nitrogen source and NaCl (1–3 g l−1) is added to the
fermentation growth. Strains have now been isolated which give product concentrations of 15 g l−1.
17.15.5.3 Tetracycline
Tetracyclines are broad spectrum antibiotics effective against gram-positive and gram-negative bacteria as well as
ricketsias, mycoplasmas, Leptospiras, Spirochetes and Chlamydias. They have a naphthalene ring system and act as inhibi-
tors of protein synthesis by preventing the binding of the amino acyl tRNAs to the ribosomal A-site. Chlortetracycline
was the first antibiotic of this group to be isolated. Tetracycline was initially synthesized from chlortetracycline but later
was found to be produced directly by Streptomyces viridifaciens. At least 20 different Streptomycetes are now known to
produce a mixture of different tetracyclines. Many mutant strains have now been developed secreting high amounts of
tetracycline. The tetracycline biosynthetic pathway requires 72 intermediate products, all of which have not yet been
characterized.
In the 1980s, however, it had become apparent that E. coli could not be used to produce some proteins particularly
requiring post-translational modification for biological activity and some other complex proteins containing multiple
disulfide bonds. In addition, many proteins expressed in E. coli were accumulated intracellularly in the form of insol-
uble, inactive inclusion bodies, from which biological activity must be recovered by complicated and costly denatur-
ation and refolding process. During the mid-1980s, these drawbacks of E. coli led to the development of animal/plant
cell expression systems. However, although such systems have proved capable of producing authentic active proteins,
they have often failed to develop the simple and efficient, high-yield, low-cost methods for the production of proteins
in large amount. In contrast, recent advances in protein refolding, translocation and roles of molecular chaperons and
foldases have made possible the design of recombinant E. coli strains that accumulate proteins in soluble form, secrete
proteins into periplasm, export proteins into culture medium and direct proteins to the outer membrane of the cell for
surface display.
Table 17-3 Techniques for improving the heterologous protein production in E. coli
Host strain Choice of host strain impacts expression
Plasmid copy number Gene dosage, as manipulated through plasmid copy number, affects expression.
Selection antibiotic Choice of antibiotic resistance on the expression plasmid can influence heterologous protein
expression.
Promoter Strong/weak, inducible/constitutive; promoter regulation is a major influence on protein expression,
which is also affected by relative orientation and strength of promoters on the plasmid.
Transcription termination Effectiveness and spacing of transcription terminators affect expression.
mRNA stability The stability of the mRNA impacts yield. Secondary structure, especially at the 5′ end of the message,
often plays a critical role.
Translation signal The ribosome-binding site affects the level of ribosome loading and clearance, and hence expression.
Secondary structure at the 5′ end of the message can affect the accessibility of the ribosome-binding site.
Codon usage Utilizing the optimal E. coli codons often improves yield.
Temperature Temperature has a pronounced effect on protein folding and stability. Lower temperature frequently
improves soluble yields.
Growth conditions/media Growth conditions, oxygen levels, growth rate, carbon source and fermenter configuration affect yield.
Fusion proteins Fusing heterologous proteins to peptide that are typically highly soluble in E. coli often improve yield
of soluble product.
strain. Streptomyces are also considered to be potentially suitable for the development of efficient secretion systems. The
reported yields are still below cost-effective ranges in most cases. The comparatively low yields demonstrate that a lot of
fundamental research is still necessary to render Streptomyces systems competitive. However a recent study comparing
recombinant alpha-amylase yields demonstrated that both the final yields and enzyme activity were considerably higher
in S. lividans in comparison to the periplasmic production in E. coli. High-level heterologous expression and secretion of
two major antigenic proteins of M. tuberculosis has been achieved in S. lividans.
Corynebacterium and related bacteria: These are a group of pleomorphic asporogenous gram-positive bacteria
found in a wide variety of ecological niches. Their major industrial importance is in the production of various amino
acids like lysine, glutamic acid, threonine, etc. Gene manipulations in these industrially important bacteria were ini-
tiated by the discovery of their indigenous small cryptic plasmids used for vector construction. Mukherjee et al. con-
structed a series of cloning vectors based on pBL1 replicon. Cloning vectors harboring hygromycin and chloramphenicol
resistance were developed by Cadenas et al. Reports on the efficient expression of foreign genes are also common. Ovine
gamma interferon was expressed as a GST fusion protein in two bacterial hosts, E. coli and C. glutamicum. Other foreign
genes successfully expressed in C. glutamicum include subtilisin of B. subtilis and basic protease of Dichelobacter nodosus.
The gene encoding fibronectin-binding protein 85A of M. tuberculosis has also been expressed in C. glutamicum using
pBL1 replicon-based vectors containing tac promoter of E. coli or endogenous cspB gene promoter. Expression of GFPUV
under the control of positively regulated araBAD promoter has been reported by Ben-Samoun et al. This study estab-
lished that C.glutamicum RNA polymerase is activated by the E. coli positive regulator of transcription AraC. Cloning
and expression of xylanase (xys1>) and the cellulase gene (celA1) from Streptomyces halstedii JM8 has been reported in B.
lactofermentum. The broad host range plasmid pEP2 was used to express the protective protease gene from D. nodosus,
causing ovine footrot in C. pseudotuberculosis. Recently, a series of gene fusion vectors have been constructed for expres-
sion of recombinant proteins in corynebacteria.
Yeast expression systems: These are unique as they offer the advantages of both being a microorganism and a
eukaryote. Its basic advantage is that it can be cultivated like bacteria on defined media, while it has the protein pro-
cessing machinery of a eukaryotic cell. Unlike E. coli, yeast provides advanced protein folding pathways for heterologous
proteins, and with yeast, signal sequences secrete the correctly folded and processed proteins. Therefore functional and
fully folded heterologous proteins can be secreted into the culture media. Unlike mammalian expression systems, yeast
can be rapidly grown on simple growth media. For the expression of clinically and industrially important proteins, yeast
is an attractive option as industrial scale fermentation technology is now widely used. Whole antibodies and antibody
fragments have been expressed using this system. The binding activity of whole antibody and Fab secreted from yeast was
found to be similar to that of their counterparts derived from lymphoid cells. Single-chain antibodies have also been suc-
cessfully expressed in yeast systems, for example, an anti-fluorescein scFv has been produced in Schizosaccharomyces pombe
and anti-recombinant human leukemia inhibitory factor scFv has been expressed in Pichia pastoris. scFv proteins which
are produced as insoluble inclusion bodies in E. coli are often soluble when expressed in yeast. In addition, the degrada-
tion of heterologous proteins, often a problem in E. coli, is usually reduced in yeast. Yeast systems include Saccharomyces
cerevisiae, P. pastoris, Hansenula polymorpha and Pichia methanolica. S. cerevisiae is a genetically well-characterized yeast
whose complete gene sequence is known. It is a commercially important strain, because it can be cultivated like bacteria
on defined media. The cultivation time for S. cerevisiae is also shorter compared to other eukaryotic cells. Recent studies
with S. cerevisiae have explored both the strong cell cycle dependence and metabolic burden of heterologous protein
secretion in these cultures. These studies provide interesting points for comparison between microbial and mammalian
systems, as well as an improved fundamental understanding of the use of yeasts for recombinant protein production.
Pichia pastoris, Hansenula polymorpha, Candida boidinii and Pichia methanolica are methylotrophic yeasts. Pichia pastoris
has developed into a highly successful system for the production of a variety of heterologous proteins. The increasing
popularity of this particular expression system can be attributed to several factors, most importantly: (1) to the presence
of strong inducible promoters (AOX1) as well as constitutive promoters (like GAP) which allow over-expression of
recombinant proteins; (2) the simplicity of techniques needed for the molecular genetic manipulation of P. pastoris and
their similarity to those of S. cerevisiae; (3) the ability of P. pastoris to produce foreign proteins at high levels, either
intracellularly or extracellularly; (4) the capability of performing many eukaryotic post-translational modifications, such
as glycosylation, disulfide bond formation and proteolytic processing and (5) the availability of the expression system as
a commercially available kit.
The optimal process conditions for human chymotrypsin B production in P. pastoris resulted in product levels of 480
mg L−1 in the broth with a biomass concentration of 150 g L−1. Alternatively, a continuous culture technique increased
the volumetric productivity by fivefold compared to fed batch culture. An insulin precursor was also expressed with
yields of 1.5 g L−1. By using more complex feeding strategies for high-cell-density cultures of P. pastoris, recombinant scFv
830 • Chapter 17/Bioprocess Engineering and Technology
antibody fragments have been produced in large amounts. Very high yields in the range of few grams per liter have been
reported for a range of proteins.
Insect cells/Baculovirus system: Baculovirus gene expression is a popular method for producing large quantities of
recombinant proteins in insect host cells. In most cases, post-translational processing of eukaryotic proteins expressed
in insect cells is similar to protein processing in mammalian cells. As a result, insect-cell-processed proteins have com-
parable biological activities and immunological reactivity to proteins expressed in mammalian cells. Protein yields from
baculovirus systems are higher, and costs are significantly lower than in mammalian expression systems. The baculovirus
expression system can express genes from bacteria, viruses, plants and mammals at levels from 1–500 mg L−1; most pro-
teins are expressed in the 10–100 mg L−1 range, although making predictions is difficult. The baculovirus most commonly
used to express foreign proteins is Autographa californica nuclear polyhedrosis virus. AcMNPV can be propagated in
certain insect cell lines; the virus enters the cells and replication begins approximately 6 hours post-infection (h.p.i.).
At approximately 20–48 h.p.i., transcription of nearly all genes ceases. The viral polyhedrin and p10 genes, however, are
transcribed at high rates. The polyhedrin protein is essential for propagation of the virus in its natural habitat; however,
in cell culture, polyhedrin is not needed, and its coding sequence can be replaced with a sequence for a target protein.
Hence, the powerful polyhedrin promoter can drive high-level transcription of the insert, resulting in expression of a
recombinant protein that can account for over 30% of total cellular protein (TCP).
Mammalian cell culture expression systems: Over the past 9–10 years, there has been a considerable progress in
fine tuning mammalian expression systems for high-level recombinant gene expression. Chinese hamster ovary (CHO)
cells and murine myeloma cells (NS0) have established themselves as predominant systems of choice for mamma-
lian expression. Mammalian cell cultures are the most acceptable way of producing eukaryotic proteins because the
authenticity of the protein is guaranteed, given the similarity of the host protein processing machinery. Thus proteins
like murine erythopoietin (EPO) and TPA are still produced for pharmaceutical purposes in mammalian cell cultures.
However, the cost of mammalian cell culture remains prohibitive because of high cost of media as well as extremely high
level of sterility to prevent contamination. Thus, scale up of mammalian cell cultures remains a big challenge not only
because of sterility requirements but also because of the shear sensitive and fragile nature of mammalian cells. Genetic
engineering approaches too have rationally modified the specific features of mammalian host cells and improved their
utility in recombinant protein expression. Recent developments in the use of serum-free media for cell cultivation
have helped cut down operational costs. Simultaneously, fine tuning of bioprocess parameters has helped increase the
productivity and yields of mammalian cell expression systems.
Cell-free expression system: Cell-free methods introduce a new degree of complexity because here the cells must
be first grown and then cell extracts prepared for further expression studies. However, the cell-free approach offers
many potential advantages; it allows the synthesis of proteins toxic to the host cells, the channelization of the met-
abolic resources towards product synthesis and also provides incredible flexibility in manipulating protein synthesis
and folding. The efficiency of ‘cell-free’ expression systems also continues to improve, both for the traditional systems
using cell extracts and for systems reconstituted from purified components. Kigawa and co-workers synthesized up to 6
mg L−1 of chloramphenicol acetyl transferase (CAT) using a continuous exchange cell-free system. Recently a report
on co-translational expression and folding of firefly luciferase using an E. coli cell-free translation system’ has been
described. Although currently useful for high throughput screening applications in this era of ‘proteomics’, it will be
interesting to see in coming years, to what extent these can be useful for large-scale protein production.
Heterologous expression has also been dealt with in Chapter 14, Section 14.13.
Copper leaching plants use a leaching solution (sulfate/Fe3+ solution) which carries the microbial nutrients in and
the dissolved copper out. It is sprinkled over the heap and percolates through the rock pile below where the copper-rich
liquid is collected. Using such methods large installations have been known to produce up to 200 tons of copper per day.
This process is called an indirect leaching process because the microbial action is not directly on the uranium ore but
on the iron oxidant. In the commercial process, the dissolved uranium is extracted from the leach liquor with organic
solvents like tributylphosphate.
17.16.2.1 Methods
The success of oil spill bioremediation depends on the establishment of favorable conditions in the contaminated
environment. The first condition is that bacteria with appropriate metabolic capabilities must be present. If that exists
then their rates of growth and hydrocarbon metabolism need to be maximized by supplying adequate nutrients and
maintaining appropriate pH conditions. One has to remember that heavy crude oils contain large amounts of resin and
asphaltene compounds that are less amenable to bioremediation than light weight oils which are rich in aliphatic com-
ponents. The oil surface area is also important as the growth of oil degrading microbes occur at the oil–water interface.
17.16.2.5 Conclusion
Research is currently going on to evaluate bioremediation and phytoremediation (plant assisted enhancement of oil
biodegradation) for their applicability to clean up oil spills contaminating salt marshes and freshwater wastelands. If
successful, it can be extremely cost effective and has the advantage that toxic hydrocarbons are degraded rather than
removed to another site. The main challenge is to maintain sufficient nitrogen and phosphorus levels for bioremediation
to occur. Studies show that application of dry granular fertilizer to the impact zone is probably the most cost-effective
833
17.17 Food Technology•
way to control nutrient concentrations. The Exxon Valdez spill that occurred in Alaska in 1987 formed the basis for a
major study on bioremediation through fertilizer application. Inipol (an oleophilic micro-emulsion with urea as a nitro-
gen source, lauryl phosphate as a phosphate source and oleic acid as a carbon source) and Customblen (a slow-release
fertilizer composed of calcium phosphate, ammonium phosphate, and ammonium nitrate within a polymerized vegetable
oil coating) were used. Within 2–3 weeks, oil on the surfaces of cobbled shorelines treated with Inipol and Customblen
was degraded so that these shorelines were visibly cleaner than nonbioremediated shorelines. The success of this field
demonstration programme has introduced bioremediation as a key technique in any clean up strategy which needs to be
developed for future oil spills.
Box: 17-2
Consider a slab though which heat flow is taking place (Figure 17-41).
Consider a thin section of thickness Δx at a distance x from one end (shaded portion). The heat flux entering at the LHS
is given by
qin = kA(dT / dx)x .
This is the Fourier equation for unsteady-state heat transfer in one dimension, which becomes dT/dt = a∇2T for three-
dimensions, where α = k/ρCp. This equation can be solved for different boundary conditions.
834 • Chapter 17/Bioprocess Engineering and Technology
17.17.2.1 Drying
Drying, the evaporative removal of water from moist solids or solidifi-
able solutions, is one of the earliest method of preserving foods. This
Figure 17-41 Temperature profile for unsteady- preservation takes place because biological activity (typically micro-
state heat transfer through a slab. bial growth) slows down drastically or stops when moisture levels fall
below 0.15 kg water/kg of dry solids. Hot air is the most common
medium used to remove moisture. Immediately after contact between the moist food and hot air the food surface tem-
perature adjusts till it reaches steady state which is given by the wet bulb temperature of the air. The drying rate also
becomes constant and this is referred to as the ‘constant rate drying period’. This period ends when the moisture content
of the solids falls below a certain level (called the critical moisture content). The drying rates fall rapidly and the surface
temperature of the food rises. For large food particles this period called the falling rate period takes a far longer time than
the constant rate period.
During the constant rate period the entire solid surface is saturated with water. Drying thus proceeds as if from a pool
of water and the solid has little role to play in the drying rate. As drying proceeds the rate of liquid movement to the
surface from the inside (by capillary action and diffusion) becomes slower than the rate of mass transfer from the surface.
Consequently the surface starts to dry and the drying rate starts getting controlled by the diffusion rate. Since this rate
of diffusion is concentration dependent, the drying rates fall in the falling rate period. Drying also cases shrinkage in
foods making a theoretical analysis of drying rates difficult. The drying air temperature and humidity affect food quality
in addition to affecting drying rates and thermal efficiency. Therefore air conditions are adjusted to minimize quality loss
and prevent in - process microbial growth.
Foods are air dried in many different types of equipment. These include – shallow-bed dryers where the food is moved
in perforated conveyor beds; deep-bed dryers where drying takes place in bins (mostly used for grains), spray driers which
are used to dry liquid foods and food slurries, freeze dryers used to dry beverage extracts, shrimp soup ingredients, etc.
17.17.2.2 Chilling
Chilling, where food is not frozen, is also widely used to extend the shelf life of fresh produce like meat fish or diary prod-
ucts. Temperatures close to 0 °C and humidities between 85–95% are usually recommended. Data for heat capacities
& heats of respiration are available for most fresh foods which allow computation of the refrigeration loads. Shelf life
can vary widely from 2 days to 12 months, though typically for a majority of products the shelf life is less than 1 month.
Chilling is combined with the use of controlled atmospheres with low oxygen (2–5%) & up to 5% CO2 which
increases the shelf life of apples, pears & cabbage. This technique is also used for long distance shipping of meat.
17.17.2.3 Freezing
Freezing preserves food by reducing the rate constants for chemical reactions, reducing reactant mobility and reducing
the water activity (aw). The water activity of a frozen food depends only on its temperature as long as the food is moist:
aw = 0.908 at −10 °C; 0.824 at −20 °C, and 0.748 at −30 °C.
Foods are frozen using a cold stream of refrigerated air, direct contacting with brine-ice mixtures, contact with inert
evaporating refrigerant or cryogenic gas or by compressing them between refrigerated plates. Most water in moist foods
will freeze but water bound to food solutes or solids in nonsolvent form will not freeze even at very low temperatures.
Food solutes also depress food freezing points. Equilibrium initial freezing points usually lie between −1 and −2 °C and
supercooling of 5 to 6 °C is required before ice nucleates and freezing starts. As freezing proceeds solutes concentrate in
the residual nonfrozen solution further reducing the freezing temperature. Freezing of water in foods thus takes place over
a range of temperatures. If freezing is slow, water gets transferred from cell interiors (especially in fruits and vegetables)
835
17.18 Enzyme Engineering•
and forms extracellular ice before nucleation occurs inside cells. This adversely affects the texture of foods. Rapid
freezing allows intracellular nucleation to take place and most of the water freezes inside the cells minimizing adverse
textural affects.
Where ‘V’ is the rate of product formation or reaction velocity. Note that the shape of this function is asymptotic,
i.e. velocity increases with increase in substrate concentration when substrate concentrations are low, (like a 1st order
reaction) and velocity saturates to a maximum value Vm when substrate concentrations are high (zero order reaction)
Vmax
V
Zero order
1st order
Km S
Vmax thus represents the maximum reaction velocity for a given amount of enzyme ET
i.e. Vmax = k 2 E T when substrate is not rate-limiting.
Exercise : Prove that the reaction velocity is half of Vmax when substrate concentration = Km and v = ¾ Vm when S = 3 Km
The constant k2 is often called kcat and represents the rate of product formed per unit amount of enzyme, when
substrate is not rate limiting. It is also called the turnover number which is the number of substrate molecules converted
to product per second by a single catalytic site, with the units of s−1. Catalase for example has a turnover number of 4 ×
107 s−1. The Km terms can be understood as the affinity of the enzyme towards its substrate, with low Km values imply-
ing high affinity and vice-versa. Clearly if Km is small then ‘V’ would remain close to Vmax and not fall as the substrate
concentration declines as long as S>>Km.
The ideal enzyme would thus have a low Km and high kcat, attaching to substrate with high affinity and converting
them efficiently to product. Thus the catalytic efficiency of an enzyme is given by kcat/ Km. Remember that the substrate
has to reach the ‘active site’ of the enzyme by diffusion. Therefore the “perfect” enzyme which ‘immediately’ converts
the substrate to product would have a Kcat/Km value limited only by diffusion control. This sets the upper limit of its
value between 108 – 109 M−1s−1 depending on the micro environment of the enzyme. The most commonly used method
to estimate the kinetic parameters Vmax and Km is the Lineweaver Burke plot of 1/V vs 1/S. Since the equation
Vmax S
V=
Km + S
836 • Chapter 17/Bioprocess Engineering and Technology
can be written as
1 K m 1 1
= +
V Vmax S Vmax
the double reciprocal plot has a slope of (Km/Vmax) and an intercept of (1/Vmax)
The method of generating such a plot is to first plot the product concentration versus time for a given amount of
enzyme using different starting concentration of the substrate. We then get a plot as shown below.
S4
S3
P
S2
S1
Here S4>S3>S2>S1 and therefore the rate of product formation rate is maximum with S4. This rate of product formation
is calculated by taking tangents and calculating its slope in the initial (early time points) part of the graph. This is not
a simple exercise especially when the Km values are small. Choosing low concentrations of the starting substrate can
cause it to decline quickly so that the velocity also declines sharply with time. This introduces errors in estimating the
slope of the tangent since we need at least a few time point samples for plotting the graph. On the other hand a high
starting value of substrate concentration leads to saturation (zero order) kinetics and thus no change is observed in ‘V’
at different substrate concentrations (as long as S>>Km). For example, if Km = 1mM then all So values of 5mM and above
would result in similar values of V (which will be close to Vmax). Conversely if the velocities (reaction rates) are around
1mM min−1 then choosing low So values would exhaust the substrate within minutes before sufficient time point samples
can be collected. A partial solution to this is choosing very dilute enzyme concentration to lower reaction velocities.
A second problem is uneven spacing of the data points while plotting the Lineweaver Burk plot. Taking reciprocals of
the ‘V’ and ‘S’ values cause the data points to cluster near the origin. Thus a straight line drawn to estimate Km and
Vmax, using linear regression would be very sensitive to errors in the smaller V and S estimates. (Since the errors in the
reciprocals would be large). We therefore use the Eadie Hofstee plots of S/V versus S (see below figure) and since;
S Km + S Km S
= = +
V Vmax Vmax Vmax
S/V
0.011
0.01
0.009
0.007 0.02
0.005 0
0 0.5 1 1.5 2 2.5 0 2 4 6 8 10 12
1/S S
An even better solution is to utilize the integral version of the Michaelis Menten, kinetics where the equation
dS V S
V=− = max
dt K m + S
837
17.18 Enzyme Engineering•
can be integrated as
Km + S
−∫ dS = Vmax ∫dt
S
S0
K m ln + (S0 − S) = Vmax t
S
Given the stoichiometry between substrate and product i.e., (S0 −S) = P, we can use this equation to simulate the
product profile w.r.t time for different Km and Vmax values using either excel or MATLAB software. One useful strategy is
to independently assess Vmax by calculating ‘V’ value with increasing substrate concentrations till it saturates. Then we
can fix Vmax and simulate the product profiles with different Km values till is best matches the experimental data.
Thus with increasing substrate concentrations the plot looks like figure shown below
Here the velocity reaches a maximum which is lower than Vmax and then declines when the substrate concentration
is increased further.
Exercise: Determine the value of the maximum velocity possible in a substrate inhibited system and the substrate
concentration at which it occurs: Hint:- Differentiate the above equation to find its maxima].
Product inhibition kinetics can also be represented by a similar equation, i.e
Vm S
V=
K m + S + P2 / K I
However determining whether product inhibition is taking place requires a careful design of experiments. This is
because now the decline in ‘V’ is observed, not in the initial stage (as with substrate inhibition) but in the later stages
when product buildup takes place (see figure below)
t
838 • Chapter 17/Bioprocess Engineering and Technology
From the figure it is clear that the product formation rate declines with time and this product formation rate can be
calculated by taking tangents(to calculate dP/dt) at different time points. However this decline in dP/dt) or ‘V’ can be
due to many reasons.
1) ‘V’ declines because the product concentration builds up and inhibits the enzyme
2) ‘V’ declines because substrate levels fall and this fall is sharper when [S] is close to Km
3) ‘V’ declines because of enzyme inactivation that takes place over time.
To distinguish between these 3 possibilities we do the following experiments
P P
t t
(a) (b)
If the product profiles with ‘high’ substrate and different enzyme concentrations is like ‘a’ above then clearly it
is product inhibited, since after reaching a certain product concentration (at different times due to different initial
velocities) the product does not increase further. However in ‘b’ above product concentrations do not increase after a
certain time has elapsed pointing to enzyme deactivation as the cause for ‘V’ going to zero. To ensure that substrate is
not rate limiting in the above experiments we need to perform the experiment alluded to earlier in the text, where ‘low
concentrations of enzyme are used and substrate concentrations are increased continuously till no significant increase in
velocity is observed (implying zero order kinetics w.r.t substrate).
Biochemists often report specific activities of enzymes as moles of product formed per unit time. This implicitly
assumes a constant velocity during the time period of measurement and zeroth order dependence on substrate concentra-
tion. As is clear from the above discussion, any change in starting substrate concentration or even the amount of enzyme
used can change these assumptions and introduce non-linear rates of product formation. One of the reasons why specific
activity measurements of enzymes like cellulases and xylanases have significant variations is precisely because of these
non-linear affects. Thus with a crystalline insoluble substrate like cellulose, Km values tend to be high and the velocity
is critically dependent of initial substrate concentrations.
For industrial uses therefore enzymes need to be characterized much more thoroughly than simple specific activity
measurements.
Diffusion processes play an important role in the design of such reactors. In the simplest case we consider enzymes
immobilized on the surface of solid matrices. If the bulk concentrations of substrate is So we can model this diffusion
process by proposing a thin film on the surface of the matrix through which the substrate diffuses to reach the enzyme,
which ‘sees’ a lower substrate concentration ‘S’. The rate of diffusion is given by
N s = k s (So − S)
where S and So are substrate concentrations at the interface and in bulk fluid respectively, and Ks is mass-transfer coeffi-
cient. At Steady State it matches the enzyme reaction rate. Thus
Vmax S
k s (So − S) = V =
Km + S
This effectively reduces the number of parameters (or variables) from 4 to 2. Solving we get
(1 − x ) = x
Da k +x
where x lies between 0 and 1.
Da is also called the Damkohler number and represents the ratio of the maximum enzymatic reaction rate to the
maximum diffusion rate. Thus low values of Da imply that the system is reaction rate limited (since reaction rate is low
compared to the rate of diffusion). Conversely a diffusion rate limited system would have high Da values. Note that the
velocity of the reaction no longer exhibits a Michaelis Menten dependence on bulk substrate concentration. We rather
define an effectiveness factor η as
observed reaction rate
h=
reaction rate if there was no diffusion ratee limitation
Two dimensionless numbers are defined; the Thiele modulus φ which is given by
R Vmax / K m S
f= and b = o
3 Des Km
where Des is effective diffusion coefficient for substrate and β is the magnitude of saturation. It is observed that the
effectiveness factor η is a function of these two dimensionless numbers, where values of φ < 0.3 gives η values close to 1
implying that the system is controlled by the enzymatic reaction rate while values of φ > 3 makes η inversely proportional
to ϕ implying that the system is controlled by diffusion.
Estimation of which factor is rate controlling can be done by choosing porous particles of different sizes. Since large
particles will give large φ the system will get diffusion controlled while small particles will be controlled by the enzymatic
reaction rate. Simultaneously we can vary the bulk concentration So and measure the corresponding reaction velocities.
Given that for large So the reaction is zero order with respect to substrate we can calculate Vm when there is no diffusion
control (for small particles). This data can then be used for measuring the change in V when the particles are larger or
when So is small to determine the effect of diffusional limitation
The above analysis ignores the effect of film diffusion and in a realistic situation we need to assess the relative values
of the film mass transfer, diffusional rates within the pore and reaction rate to determine which of these is the rate lim-
iting factor in enzyme kinetics.
Keywords
Aiba. S., A.E. Humphrey and N.F. Millis (1973) Biochemical Stanbury, P.F. and A. Whitaker (1989) Principles of
Engineering, University of Tokyo Press, Tokyo. Fermentation Technology, Pergamon Press, Oxford.
Atkinson, B. (1982) Biochemical Reactors, Pion Ltd., Nielsen, J. and J. Villadsen (Eds.) (1994) Bioreaction
London. Engineering Principles, Plenum Press, New York.
Baily, J.E. and D.F. Ollis (1986) Biochemical Engineering Schuler, M.L. (Ed.) (1989) Chemical Engineering Problems
Fundamentals, McGraw-Hill Book, New York. in Biotechnology, AICHE, New York.
Enfors-Haggström (2000) Bioprocess Technology: Lee, J.M. (1991) Biochemical Engineering, Prentice Hall,
Fundamentals and Applications, KTH, Stockholm. Englewood Cliffs, NJ.
Jackson, A.T. (1991) Process Engineering in Biotechnology, Vieth, W.F. (1994) Bioprocess Engineering–Kinetics, Mass
Prentice Hall, Englewood Cliffs, NJ. Transport, Reactors and Gene Expression, John Wiley &
Shuler, M.L. and F. Kargi (2002) Bioprocess Engineering: Sons, New York.
Basic Concepts, Prentice Hall, Upper Saddle River, NJ.
Model Questions
from second reactor. Would the answer change if we If the same broth is now to be filtered in an industry
used only one reactor of (2+2) = 4l. using a plate and frame filter of 10 m2 surface area till
the cake is 3 cm thick. Determine volume processed
F0.5 lh1
sF 20 gl1 and time required per batch. [Hint: Plot reciprocal of
F1lh1 filtration rate vs filtrate volume to determine filtration
sF 10 gl1 constants.]
7. Distillation of a two-component mixture (A and B) is
carried out in a batch mode. If α = 1.5 and the initial
mixture has a 50:50 ratio of both components, deter-
mine the composition of the distillate when 60% of
the distillate has been collected. For the same mix-
V2 l V2 l ture determine graphically the minimum number of
plates required (at total reflux) to get a distillate con-
[Hint: Use the material balance equations to derive the centration of 95% A, 5% B and a bottoms product
outlet biomass and substrate concentration from the of 90% B and 10% A in a distillation column. [Hint:
first reactor; Use biomass balance on second reactor, Plot the equilibrium curve, since the operating line
overall balance on substrate and the kinetic equation at total reflux corresponds to the 45 °C diagonal the
to get biomass and substrate at the outlet of the second minimum number of plates can be easily determined.]
reactor. Use product balance to determine product 8. A 100 l reactor gets sterilized (as per design c alculations)
concentration.] by holding it for 5 min at 121 °C. If this sterilization is
to be scaled up to 50,000 l what should be the hold-
3. A waste treatment external recycle reactor has to run ing time. (State your assumption clearly). Given are
so that the outlet biomass concentration xout should not K121 °C = 2.54 min−1 ∇121 °C = 12.5 (Richards). [Hint:
/ s = 0.5 g/g, ms = 0.1
exceed 5 gl−1. Feed, SF = 20 gl−1, Yxtrue Note that the ∇ factor increases since the initial cell
h−1, µm = 1.0 h−1, Ks = 20 mg l−1 and D = 0.5 h−1. Determine count increases by a factor of 500.]
the fraction of cells to be recycled if the cells are con- 9. An acid-producing fermentation is controlled at pH 7
centrated 10-fold in the sedimentation tank. (Section by a proportional controller with a proportional band
17.5.10) (PB) = 2. The pH meter reads 6.7 continuously. Why?
4. Explain how an enrichment strategy using continuous Would this reading change if PB = 1? Can the PB be
culture techniques is better than serial inoculations. reduced further? Discuss. (Section 17.9.2)
How would this help in screening for desirable charac- 10. In a dynamic gasing out method air is shut off and the
teris other than the yield of product. (Figure 17-34 and DO reading recorded at 20 intervals. These are given
accompanying text) below.
5. How would the choice of media differ when industrial
alcohol is being produced in comparison to production DO (%) 80, 59, 37, 17 (after 10 s), 9 (air put on).
of recombinant therapeutic protein. Discuss the sim- DO (%) 39, 53, 63, 70, 76, 78.
plex algorithm for determining optimum media con- Determine KLa. If air is bubbled at 1vvm what outlet
centrations. (Section 17.11 and also references at the oxygen concentration would you get.
end of the chapter) Hint: Plot ln/ (C / L − CL )min / (C / L − CL ) vs. time to
6. Filtration carried out using a plate and frame filter of cul- get KLa.]
ture fluid gave the following data. Filter area = 250 cm2, 11. Given the KLa, α(P/V)0.8 (Rs)0.3, do a scale-up on
cell density = 7 g l−1 in culture fluid, dried cake den- the basis of constant KLa from a 5 l to 10,000 l
sity = 1.5 g cm−3, ΔP = 2atm (constant). fermentor given that the air flow rate is reduced
from 1 vvm to 0.5 vvm. By what ratio would the rpm
Volume of filtrate(l) Time (s)
change?(Section 17.8)
0.2 1.8 12. Glucose and yeast extract (YE) are to be optimized
0.4 4.2 using Box–Wilson design. The range of concentration
0.6 7.5 to be used is 0.5 to 2% and 0.1 to 1% for glucose and
0.8 11.2 YE respectively. What are the concentration to be
1.0 15.4 used in the individual experiments. IF high glucose
1.2 20.5
concentration repress product synthesis while high YE
1.4 26.7
1.6 33.4
concentration favour product formation how would
1.8 41.0 the contour plot look. (Section 17.11.1.2)
2.0 48.8 13. Design a sterilization protocol for a 10,000 l
2.2 57.7 fermentation media containing 103 cells per ml ini-
tially. To prevent nutrient loss at least 50% of the
842 • Chapter 17/Bioprocess Engineering and Technology
sterilization should be during the holding period 20. A recycle reactor is run where the outgoing cells from
at 121 °C. Given at T = 121 °C, k = 2.538 and V121 the reactor are fed back in a highly concentrated form
°C = 12.549 (Richards). [Hint: Assume equal rates of to the reactor. Given mm = 1.0 h−1 of KS = 20 mg l−1,
heating and cooling and therefore equal contribution D = 0.5 h−1 of So = 10 g l−1Yx/s = 0.5 g/g the following
to the ∇ factor.] results are obtained when different percentages of the
14. A fermentation running at 37 °C is to be maintained outgoing cells from the reactor are fed back.
at the temperature using cooling water at 12 °C. If
80% of the heat generated is due to metabolic energy % fed back Product concentration (gl−1)
(@ 5kcal/l/h). Determine the cooling surface area 10% 2.25
requirement per m3 of fermentor volume. Given: Rise 20% 2.51
in cooling water temperature = 5 °C, Uoverall = 1000 w/ 30% 2.86
m2 K. [Hint: See text for worked example.] 40% 3.33
15. Yeast cells. (CH1.8 O0.5 NO2) are grown anaerobically
to produce ethanol. The carbon source is glucose Determine product formation kinetics (α and β).
and the nitrogen source is glutamate (C5H9O4N). If [Hint: Given Ks is small the outlet biomass
the true yield ( Yxtrue concentration would not change (overall balance)
/ s ) on glucose under these circum-
stances is 0.1 g/g and ms = 0.1 g/gh. Calculate the prod- while the µ would change since D is constant and
uct yield Yp/x at a specific growth rate of 0.1 h−1. [Hint: recycle is changing. Use the product balance equation
Calculate Yxobs (αμ + β)x = Dp to graphically determine the result.]
/ s and then use stoichiometry.]
16. E. coli cells (mm = 1.0 h−1, Ks = 10 mg l−1, Yx/s = 0.5 g/g) 21. A plug flow reactor has to be designed to treat liquid
are grown in fed-batch culture using constant feed @ waste comprising 5% starch at the rate of 106 l/year.
200 ml h−1 Sfeed = 10 gl−1 after growing the cells in batch Given outlet starch concentration has to be ≤0.5%,
culture Vo = 2l, So = 10/g/l). The fed batch is done for determine the reactor size if the organism has a
10 h. Assuming negligible product concentration at mmax = 0.5 h−1 and Ks = 100 mgl−1. Assume that 5% of
the beginning of feed calculate final product concen- the outlet biomass is recycled with the inlet feed.
tration given α = 0.2 and β = 0.1 (state your assump- If a CSTR were used would the size change? Given
tions clearly). [Hint: Since Ks is small assume all the Yx/s = 0.5 g/g.(Section 17.5.3)
feed is consumed leading to linear growth profile. 22. Pichia pastoris is grown on a mixture of methanol
Hence set up the integration to determine product YE (which comprises 50% protein having an approx-
concentration.] imate formula C5H9O4N and the remaining as min-
17. Cooling water enters at 15 °C and leaves at 25 °C erals and vitamins). If only biomass is formed what
thereby removing the metabolic heat from a CSTR yield would you expect in terms of gms of biomass per
being run at 37 °C. If m = 0.2 h−1 and x = 25 gl −1 gram of methanol. Also what would be the stoichio-
calculate cooling coil surface area per m3 of fermen- metric balance between methanol and YE. What is
tor volume. Given Ui = 1000 w m−2 k−1 the carbon the O2 consumption per gm of biomass formed. (Hint:
source is glucose, nitrogen source is ammonia and the Assume hH = 0.6 or Yx / s = 0.6 .
enthalpy efficiency = 0.6. (Section 17.6) 23. An anaerobic culture of Zymomanas mobilis is run on
18. Derive the expression for KLa using the two- film glucose and NH3. If Yxtrue / s = 0.2 C-mole/C-mole and
theory of mass transfer and Henry’s law. Given dis- ms = 0.2 C-mole/C-mole h, determine outlet biomass,
solved oxygen (DO) concentration at saturation substrate and ethanol concentration at a dilution rate
(100%) = 8 mg l−1 and the fermentor is to run at 40% of 0 h−1. Given S0 = 10 g l−1, m m = 0.5 h −1 K s = 200 mgl −1
DO. Calculate the maximum biomass concentration (Hint: Determine Yxobs / s and use the stoichiometric rela-
it can attain at a specific growth rate of 0.3 h−1 if tionship to determine Yp/s)
KLa = 100 h−1. Given 24. What kind of yield coefficient (Yx/s) would you predict
for cells growing on (a) glucose; (b) methanol. State
(a) carbon source is glucose; clearly the principle you would use in each case. What
(b) carbon source is methanol. values of Yx/o would you get. Assume that no product
(State your assumptions. Hint: Note that Yx/o is is formed and NH3 is the nitrogen source. (Hint:
d ifferent for different carbon sources.) Determine whether the system is enthalpy limited or
19. Scale up is done keeping KLa constant from 1 l to carbon limited.)
106 l using the correlation KLa α(P/V)0.7 (Vs)0.2. If the 25. Determine the reactor volume required to produce
air flow rate/unit vol. (vvm) is kept constant during a product X @ 1 ton/yr. Given that the product is
scale-up, how does the power consumed/unit volume growth-associated with Yp / x = 0.2g/g and the batch
change. (Section 17.8) reactor is run with S0 = 50 gl −1 to a point where
843
Model Questions•
99% of the substrate is consumed; also given that concentration and the value of Yx/s to get inlet sub-
Yx / s = 0.5g/g and 1 year has 300 working days; the strate concentration. Determine OUR and hence the
specific growth rate is 0.5 h −1 and K s is 20 mgl −1. State outlet oxygen concentration.)
your assumptions clearly. 31. Yeast cells grown anaerobically with NH3 as the
nitrogen source produce 0.1 g cells/g of glucose.
(Determine the product concentration at the end of What values of Yp/s do you expect. (Hint: Use the
the run. Also determine the time required for one run stoichiometric equation.)
by noting that the substrate concentration at the end 32. If Yxtrue
/s is 0.5 g/g and ms = 0.2, design an external
of the run is still higher than Ks.) feedback CSTR which gives a Yx/s of 0.3 g/g at a
26. A CSTR is run and the following values are obtained dilution rate of 0.5 h −1 . Given the concentration
by the researcher assuming that it is an ideal well- factor in the external sedimentation tank is equal to
mixed reactor, m m = 1.0 h −1 , K s = 20 mgl −1 Yx / s = 0.5g/g 10, determine the recycle ratio (R). Also determine
ms = 0. However it turns out that due to poor mixing Yo/x and heat dissipated per gm of cells formed. (Hint:
the concentration of cells in the bottom half of the See worked example in text.)
reactor is twice the top half from where the outlet is 33. Cells growing on glucose and glutamic acid as nitrogen
connected. How will the above parameters be reevalu- source has a Yx / s = 0.7 g/g . No product is formed.
ated taking this nonideality into consideration. (Hint: Determine Yo/x and ηH. (Hint: Use the material and
Set up the material balance equations as in the case of energy balance equation.)
internal recycle.) 34. Determine KLa if the outlet air contains O2 at a partial
pressure of 0.16 atm. Air flow rate = 0.5 vvm, C∗ = 8 mgl −1
at 100% saturation and C (steady state) = 50%. (Hint:
Use the gas balance equation to determine OTR.)
35. Scale up a process from 1 l to 10,000 l with constant
KLa a(P / V)0.8 Vs0.3 . Given that the initial air flow
rate = 1vvm and N = 600 rpm and final air flow
X rate = 0.5 vvm, determine final rpm. What would
be the final rpm if scale-up is done on the basis of
constant (P/V). (Hint: See worked example in text.)
2X 36. A 5l CSTR is to be operated at a dilution rate (D)
of 0.3h−1 using inlet glucose concentration (S0) = 10
27. What is meant by quasi-steady-state in fed-batch reac- gl−1 and ammonia as nitrogen source mm = 1 h−1 and
tors. Derive the expression μ = D for fed-batch and Ks = 50 mgl−1. Cooling water is available at 15 °C with
explain. (Section 17.5.11) a flow rate of 100 ml min−1. Calculate the length of
28. A CSTR is run at a dilution rate of 0.2 h −1 where the tubing required to maintain the fermentor at its opti-
outlet biomass concentration is 20 gl −1. The substrate mum temperature of 37°C by removing the heat of
is glucose, ammonia is the nitrogen source. The true fermentation. Clearly state your assumption. Given
−1
/ s = 0.5 g/g, while m s = 0.2 h . Determine
yield is Yxtrue that tube is thin copper of diameter = 5 mm, m = 1
the surface area of cooling required per liter of reac- cp (0.001kg/ms) of water, k = 0.15 cal/sm2 (°C/m),
tor volume given that reactor operating temperature Cp = 1 cal/g °C(1000 cal/kg°C), Nu = 0.023 Re0.8(Pr)1/3
= 37 ° C , cooling water inlet = 10 ° C, outlet = 15 ° C, (for internal heat transfer coefficient) and
U = 100 wm −2 k −1 and the heat removed is essentially h0 = 500 cal / sm 2 °C . (Hint: Determine hi from Nu,
the heat produced due to fermentation. (Section 17.6) hence determine U calculate qmet and therefore the
29. Determine Yo/x for cells growing on (a) glucose and rise in temperature of cooling water, calculate LMTD
(b) methanol if Yx / s = 0.5 g/g. What is the ηH in each and hence area of cooling required.)
case. What is the heat generated (D)? (Hint: Use the 37. A two-stage CSTR as shown in figure is used to grow
material and energy balance equation.) recombinant E. coli. The cells are grown in the first
30. In the previous question if the cells were growing in a stage and induced in the second with addition of
CSTR (D = 0.5 h −1) and the outlet biomass concentra- IPTG (small amount). Recombinant protein pro-
tion x = 1.5 mgl −1. Determine outlet substrate concen- duction is growth associated with α = 0.3. The μm
tration ( s ) and outlet oxygen concentration. Given of induced cells is half the μm of noninduced cells.
m m = 0.8 h −1 K s = 50 mgl −1, aeration rate = 1vvm , Calculate outlet product concentration given that it
what is the inlet substrate concentration (S0). (Hint: is efficiently secreted out into the medium. (Hint: See
Use the kinetic equation to get the outlet substrate question 2.)
844 • Chapter 17/Bioprocess Engineering and Technology