Skip to main content

Scale-Up of Bioprocesses

  • Chapter
  • First Online:
Bioreaction Engineering Principles

Abstract

The previous chapters in this book have dealt with stoichiometric, thermodynamic and kinetic analysis of bioreactions, as well as the operation of small-scale, “ideal,” bioreactors. These subjects constitute the basis for exploitation of microorganisms in fermentation processes. The ultimate goal for process development is, however, the realization of large-scale commercial production. The basis for a successful scale-up is to follow the advice of H. Baekeland, the inventor of Bakelite: “Commit your blunders on a small scale and make your profits on a large scale.” This is, however, not always easy. Many different engineering tools need to be applied (Leib et al. 2001), and the final scaled-up process will necessarily be a delicate compromise between inherently conflicting desirable options. Furthermore, even if the best engineering judgment is used, there are sometimes surprises in the final process which were difficult to anticipate from the lab-scale experiments.

This is a preview of subscription content, log in via an institution to check access.

Access this chapter

eBook
USD 16.99
Price excludes VAT (USA)
  • Available as EPUB and PDF
  • Read on any device
  • Instant download
  • Own it forever
Softcover Book
USD 109.99
Price excludes VAT (USA)
  • Compact, lightweight edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info
Hardcover Book
USD 109.99
Price excludes VAT (USA)
  • Durable hardcover edition
  • Dispatched in 3 to 5 business days
  • Free shipping worldwide - see info

Tax calculation will be finalised at checkout

Purchases are for personal use only

Institutional subscriptions

Notes

  1. 1.

    Andrei N. Kolmogorov (1903–1987) was a famous Russian mathematician who contributed significantly to statistical mechanics, stochastic processes, fluid mechanics, and nonlinear dynamics. His papers from 1941 on turbulent flow have been of fundamental importance for the development of this scientific field.

  2. 2.

    Claude Louis Navier and George Gabriel Stokes probably never met. Stokes (1819–1903) matriculated (Pembroke College, Cambridge) in 1837, the year after the death of Navier (1785–1836), but both contributed greatly to the development of Fluid Dynamics. In 1822, Navier who had little knowledge of the shear stress concept, i.e., on motion with friction, still developed the correct form of the equation. Stokes who held the Lucasian professorship at Cambridge from 1849 to his death put Fluid Dynamics on the right physical foundation in a series of papers from the 1840s. Both Navier and Stokes are to be counted among the giants of Applied Science and they were experts in many fields of Engineering, e.g., bridge-building.

References

  • Amanullah, A., McFarlane, C.M., Emery, A.N., and Nienow, A.W. (2001). A scale-down model to simulate spatial pH variations in large-scale bioreactors. Biotechnol. Bioeng. 73:390–399.

    Article  CAS  Google Scholar 

  • Andric, P. (2010). Reactor design for improved enzymatic hydrolysis of lignocellulose. PhD thesis, DTU (DK).

    Google Scholar 

  • Bird, R.B., Stewart, W.E., and Lightfoot, E.N. (2002). Transport phenomena, 2nd ed., John Wiley & Sons, New York.

    Google Scholar 

  • Brown, D.A.R., Jones, P.N., and Middleton, J.C. (2004) Experimental methods, part A: Measurement tools and techniques for mixing and flow-visualization studies. In Handbook of Industrial Mixing (E.L. Paul, V.A Atiemo-Obeng and S.M. Kresta, eds.) Wiley Interscience 145–202.

    Google Scholar 

  • Bylund, F., Castan, A., Mikkola, R., Veide, A., and Larsson, G. (2000). Influence of scale-up on the quality of recombinant human growth hormone. Biotechnol. Bioeng. 69:119–128.

    Article  CAS  Google Scholar 

  • Cooke, M. Middleton, J.C., and Bush, J. (1988). Proceedings 2nd Int. Conf. Bioreactor fluid dynamics, BHR group 37–64.

    Google Scholar 

  • Crank, J. (1956). The Mathematics of Diffusion, 1st edition Oxford (UK).

    Google Scholar 

  • Delaplace, G., Bouvier, L., Moreau, A., Guérin, R., and Leuliet, J.C. (2004). Determination of mixing time by colorimetric diagnosis-application to a new mixing system. Exp. Fluids, 36: 437–443.

    Article  CAS  Google Scholar 

  • Dominguez, J., Abreu, A.M., McCalla, R., Borroto, J. Ortueta, M., and Perez, E. (1999). Mixing characterization in batch reactors using the radiotracer technique. J. Radioanal.& Nucl. Chem, 241:337–340.

    Google Scholar 

  • Einsele, A., Ristroph, D.L., and Humphrey, A.E. (1978) Mixing times and glucose uptake measured with a fluorometer. Biotechnol. Bioeng. 20:1487–1492.

    Article  CAS  Google Scholar 

  • Enfors, S.O. (ed) 2001. Physiological responses to mixing in large scale reactors. J. Biotechnol, 85: 175185.

    Google Scholar 

  • George, S., Larsson G., Olsson, K., and Enfors, S.-E. (1998) Comparison of the Baker’s yeast process performance in laboratory and production scale. Bioproc. Eng. 18:135–142.

    Article  CAS  Google Scholar 

  • Grenville, R.K., and Tilton, J.N. (1996). A new theory improves the correlation of blend time data from turbulent jet mixed vessels, Chem Eng Research Design, 74:390–396.

    CAS  Google Scholar 

  • Hua, L., Nordkvist, M., Nielsen, P.M., and Villadsen, J. (2006). Scale-up of enzymatic production of lactobionic acid using the Rotary Jet Head System, Biotechnol. Bioeng. 97:842–849.

    Article  Google Scholar 

  • Hummer, J.S. (2004). A method and a process plant for treating a batch of liquids. European Patent 1,324,818.

    Google Scholar 

  • Johnson, A.T. (1999). Biological process engineering, John Wiley & Sons, New York.

    Google Scholar 

  • Kold, D. (2010). Study of mass transfer in viscous fermentation – using a Rotating Jet Head mixing system. PhD thesis DTU (DK).

    Google Scholar 

  • Leib, T.M., Pereira, C.J., and Villadsen, J. (2001). Bioreactors, a chemical engineering perspective. Chem. Eng. Sci. 56:5485–5497.

    Article  CAS  Google Scholar 

  • Masiuk, S. (2000). Mixing time for reciprocating plate agitator with flapping blades. Chem. Eng. J. 79:23–30.

    Article  CAS  Google Scholar 

  • McFarlane, C.M., and Nienow, A.W. (1995). Studies of high solidity ratio hydrofoil impellers for aerated bioreactors, part I and II, Biotechnol. Prog. 11:601–607 and 608–618.

    Google Scholar 

  • Melton, L.A., Lipp, C.W., Spralding, R.W., and Paulson, K.A. (2002). DISMT determination of mixing time through color changes. Chem. Eng. Commun. 189:322338.

    Article  CAS  Google Scholar 

  • Metzner,A.B., and Otto, R.E. (1957). Agitation of non-Newtonian fluids, AIChE J, 3:310.

    Article  CAS  Google Scholar 

  • Nienow, A.W. (1997). On impeller circulation and mixing effectiveness in the turbulent flow regime, Chem. Eng. Sci. 15:2557–2565.

    Google Scholar 

  • Nienow, A.W. (2010). Stirred tank bioreactors, and Scale-up and scale down of stirred bioreactors (in Comprehensive Bioprocess Engineering, Berovic, M. and Enfors S.O.(editors), University of Ljubljana (Slo), 135–159 and 179–199.

    Google Scholar 

  • Nienow, A.W., and Lilly, M.D. (1979). Power drawn by multiple impellers in sparged vessels. Biotechnol. Bioeng. 21:2341–2345.

    Article  Google Scholar 

  • Nordkvist, M., Grotkjær, T., Hummer, J.S., and Villadsen, J. (2003). Applying rotary jet heads for mixing and mass transfer in a forced recirculation tank reactor system, Chem. Eng. Sci. 58:3877–3890.

    Article  CAS  Google Scholar 

  • Nordkvist, M., Nielsen, P.M., and Villadsen, J. (2006). Oxidation of lactose to lactobionic acid by a Microdochium nivale carbohydrate oxidase; Kinetics and operational stability. Biotechnol. Bioeng. 97:694–707.

    Article  Google Scholar 

  • Nordkvist, M., Vognsen, M., Nienow, A.W., Villadsen, J. and Gernaey K.V. (2008). Mixing by Rotary Jet heads: Indications of the effect of rotation at turbulent and transitional flow conditions. Transactions, 6th symposium on mixing in industrial process industries, Niagara Falls (Ca) 17–21 August 2008.

    Google Scholar 

  • Oldeshue, J. (1966). Fermentation mixing scale-up techniques, Biotechnol. Bioeng. 8:3–24.

    Article  Google Scholar 

  • Oosterhuis, N.M.G., and Kossen, N.W.F. (1983) Oxygen transfer in a production scale bioreactor. Chem. Eng. Res. Dev. 61:308312.

    CAS  Google Scholar 

  • Otomo, N., Bujalski, W., and Nienow, A.W. (1995). The application of a compartment model to a vessel stirred with either dual radial or dual axial flow impellers. 1995 IChemE Research Event, Proceedings, 829–831.

    Google Scholar 

  • Otomo, N., Bujalski, W., Nienow, A.W., and Takahashi, K. (2003). A novel measurement technique for mixing time in an aerated stirred vessel. J. Chem. Eng. Japan, 36:66–74.

    Article  CAS  Google Scholar 

  • Parente, E. and Ricciardi, A. (1999). Production, recovery and purification of bacteriocins from lactic acid bacteria. Appl Microbiol. Biotechnol. 52:628–638.

    Article  CAS  Google Scholar 

  • Pedersen A.G. (1992). Characterization and modeling of bioreactors. PhD thesis, DTU (DK).

    Google Scholar 

  • Pedersen, A. G., Bundgård-Nielsen, M., Nielsen, J., Villadsen, J (1994).Characterization of mixing in stirred tank bioreactors equipped with Rushton turbines. Biotechnol. Bioeng. 44:1013–1017.

    Article  CAS  Google Scholar 

  • Picque, D., and Corrieu, G. (1988). New instrument for on-line viscosity measurement of fermentation media. Biotechnol. Bioeng. 31:19–23.

    Article  CAS  Google Scholar 

  • Revill, B.K (1992). Jet mixing (in Mixing in the Process Industries, Harnby, N., Edwards, M.F., and Nienow, A.W. (editors), Oxford, UK) 159–183.

    Google Scholar 

  • Rushton, J.H., Costich, E.W., and Everett, H.J. (1950a). Power characteristics of mixing impellers – Part I. Chem. Eng. Prog. 46:395–404.

    CAS  Google Scholar 

  • Rushton, J.H., Costich, E.W., and Everett, H.J. (1950b). Power characteristics of mixing impellers – Part II. Chem. Eng. Prog. 46:467–476.

    CAS  Google Scholar 

  • Schweder, T., Kruger, E., Xu, B., Jurgen, B., Blomsten, G., Enfors, S.-E., and Hecker, M. (1999). Monitoring of genes that respond to process-related stress in large-scale bioprocesses. Biotechnol. Bioeng. 65:151–159.

    Article  CAS  Google Scholar 

  • Sweere, A.P.J., Luyben, K.Ch.A.M., and Kossen, N.W.F. (1987). Regimen analysis and scale-down: tools to investigate the performance of bioreactors. Enz. Microb. Technol. 9:386–398.

    Google Scholar 

  • Villadsen, J. (1997). Putting structure into Chemical Engineering. Chem Eng Sci, 97: 2857–2864.

    Google Scholar 

  • Vrabel, P., van der Lans, R.G.J.M., Luyben, K.Ch.A.M., Boon, L., Nienow, A.L. (2000). Mixing in large-scale vessels stirred with multiple radial or radial and axial up-pumping impellers: modelling and measurements. Chem. Eng. Sci. 55: 5881–5896.

    Google Scholar 

Download references

Author information

Authors and Affiliations

Authors

Corresponding author

Correspondence to John Villadsen .

Problems

Problems

Problem 11.1 Scale-up without maintaining geometrical similarity

In Example 11.9 and 11.10, linear scale-up of a stirred tank bioreactor from 600 L to 60 m3 was considered. It was found that the mixing time increased by a factor of ≈3. Suppose now instead that the impeller diameter of the 60 m3 reactor is chosen to 1.4 m (i.e., the reactor geometry is changed), but the requirement of a constant P/V (113 W m−3) is maintained.

  1. (a)

    How much larger is the mixing time in the large-scale bioreactor now?

  2. (b)

    Determine the value of N that would give the same t mix as in Example 11.9 with [P/V, d s] = [113 Wm−3, 1.4 m].

  3. (c)

    Would the choice of [N, P/V, d s] in (b) increase or decrease the risk of flooding?

  4. (d)

    At a certain t mix, and for n RJH with nozzle diameter d you have calculated a power input P/V and a corresponding liquid flow v 0l for V = V 0. Show that for the same P/V, but at a different medium volume V 1 the liquid flow must be v 1l = v 0l (V 1/V 0)1/3. Use the data from Example 11.9 to calculate v l for V 1 = 3.4 m3 based on v l = 199.2 m3 h−1 that was obtained for V 0 = 60 m3.

Problem 11.2 Exchanging impellers

One drawback of the traditional Rushton impeller is that the ratio between aerated and unaerated power consumption falls rapidly with an increased aeration rate. The hydrofoil impeller typically has a lower power number, but the aerated power consumption falls less rapidly with increasing aeration rate. For a Rushton turbine and a Prochem impeller, and for stirring speeds close to 300 rpm, the ratio between aerated and unaerated power consumption is given in the figure shown below.

You are replacing a Rushton turbine in a 100 m3 reactor (T = 3 m, d s = 1 m) by a Prochem impeller. Assume that the stirring rate should be maintained the same.

  1. (a)

    What diameter of the Prochem impeller will give the same unaerated power consumption as the Rushton turbine? The values of N p for the Rushton turbine and the Prochem impeller are 5.2 and 1.5, respectively.

  2. (b)

    How much higher (approximately) will the k l a value be for the Prochem impeller at an aeration rate of 0.5 vvm under the conditions in (a), (i.e., the same unaerated power consumption is achieved)? The k l a value for the same specific power input has been found to be identical for the two impeller types. Assume a non-coalescing medium.

figure a

Problem 11.3 Design of a pilot plant bioreactor

In connection with the purchase of a new pilot plant bioreactor (41 L) to be used for penicillin fermentation, it is desired to examine whether one of the manufacturer’s standard-design bioreactors (equipped with Rushton turbines) can be used. The dimensions of this bioreactor are specified in the table below.

Aspect ratio (T/d s)

3

Tank diameter (T)

0.267 m

Stirrer diameter (d s)

0.089 m

Number of impellersa

3

Maximum stirring speed (N max)

600 rpm

  1. aThe impellers are six-bladed Rushton turbines.
  1. (a)

    Show that with a non-Newtonian medium, for which the rheology is described by a power law expression (11.26), Res is given by (1)

    $$ {{\rm Re}_{\rm{s}}} = \frac{{{\rho_{\rm{l}}}{N^{{2 - n}}}d_{\rm{s}}^{{2}}}}{{K{k^{{n - 1}}}}}, $$
    (1)

    where k = 10 is the constant used for calculation of the average shear rate in (11.32). For a medium containing, respectively, 0, 20, and 40 g L−1 biomass (of the fungus P. chrysogenum) you are required to plot Res vs. the stirring speed.

    Discuss the results.

  2. (b)

    Determine the power number with N = 600 rpm for the three cases considered in (a), and calculate the power input per unit volume when the bioreactor contains 25 L of medium. Calculate the average viscosity in the bioreactor.

  3. (c)

    For non-Newtonian media and the 41 L bioreactor, the following correlation was found:

    $$ {k_{\rm{l}}}a = 0.226 \times {10^{{ - 3}}}u_{\rm{s}}^{{0.4}}{\left( {\frac{{{P_{\rm{g}}}}}{V}} \right)^{{0.6}}}{\eta^{{ - 0.7}}}. $$
    (2)

    Can the dissolved oxygen concentration in the reactor be maintained above 30% (which for some strains is a critical level for penicillin production) when the oxygen requirement for a rapidly growing culture of P. chrysogenum is r o = 2.3 mmol of O2 (g DW)−1 h−1 Discuss how the bioreactor can be modified to satisfy the oxygen requirement.

  4. (d)

    You decide to examine the effect of increasing the stirrer diameter. Start with d s /T = 0.4. Can the critical level of dissolved oxygen concentration be maintained with this diameter ratio?

Problem 11.4 Scaled-down experiment

The pH value is often controlled using only a single pH electrode and a single point of addition of base or acid, also in large-scale bioreactors. The pH electrode is typically located in a well-mixed region, and addition of base or acid is typically made at the liquid surface. Since concentrated solutions are used, pH gradients in large-scale reactors are likely to be present.

Amanullah et al. (2001) used of a scaled-down system to study effects of pH gradients. The scaled-down system consisted of a 2 L standard stirred tank reactor, with a working volume of 1 L, equipped with two Rushton turbines (d s /d t = 0.33). To the reactor was connected a piece of tubing (L = 2.75 m, d i = 4.8 mm), through which liquid from the reactor was pumped. The pH value in the reactor was measured, and pH control was achieved by adding base to a small mixing bulb (volume about 1.5 ml) located just before the tubing (see figure below).

figure b

This system was used to experimentally simulate a three compartment reactor model, with a direct feed zone (bulb), a poorly mixed zone (tubing part), and a well-mixed zone (the reactor).

  1. (a)

    Discuss what residence time in the tubing that should be chosen to simulate a large-scale reactor (100 m3). The same residence time in the tube can be achieved by different ratios of tube volume, V tube and recirculation flow, v rec. Discuss how the values of V tube and v rec should be chosen.

    In their study, Amanullah et al. (2001) used a strain of Bacillus subtilis. This organism produces acetoin and 2,3 butanediol under oxygen-limited conditions. The formation of acetoin and butanediol is described by:

    $$ {\hbox{2 pyruvate}} \to {\hbox{C}}{{\hbox{O}}_{{2}}} + {\hbox{ acetolactate}} \to \left( {\hbox{acetolactate synthase}} \right) $$
    $$ {\hbox{Acetolactate}} \to {\hbox{C}}{{\hbox{O}}_{{2}}} + {\hbox{ Acetoin}} \to \left( {\hbox{acetolactate decarboxylase}} \right) $$
    $$ {\hbox{Acetoin }} + {\hbox{ NADH }} + { }{{\hbox{H}}^{ + }} \leftrightarrow {2},{\hbox{ 3 butane-diol}} \to \left( {\hbox{butanediol dehydrogenase}} \right) $$

    At pH values higher than 6.5 also acetate is formed.

  2. (b)

    Compare the residence time in the tube to the characteristic times for oxygen consumption and substrate consumption. Assume that the dissolved oxygen concentration in the reactor is 10% of DOT, and that the system is to be operated as a chemostat with a glucose concentration = 10 mg L−1 and a biomass concentration of 4 g L−1.

    What are your conclusions?

  3. (c)

    The experiments by Amanullah et al. were in fact made as batch cultivations, with oxygen-limited conditions also in the stirred tank reactor. A control experiment, in which base addition was made in the reactor instead of in the loop was also made. The following yields were found

    Residence time in loop (s)

    Yield of acetic acid (g g−1)

    Sum of yields of acetoin and butanediol (g g−1)

    Biomass yield (g g−1)

    (no loop)

    0

    0.30

    0.49

    30

    0.002

    0.31

    0.52

    60

    0.023

    0.29

    0.51

    120

    0.076

    0.23

    0.50

    240

    0.083

    0.22

    0.50

    120a

    0

    0.30

    0.50

    1. apH control made in the stirred tank reactor instead of in the loop

    Discuss the results.

Problem 11.5 Calculation of the true v l and the true power input for Rotating Jet Heads

In Example 11.10 the mass transfer coefficient k l a was calculated from (4), assuming that for a given Δp, v l was given by (11.37) and (11.35).

When the medium is aerated with v g /v l = n t then, for a given Δp one must use (11.36) to calculate v l rather than (11.35) which is derived for unaerated medium. v l will be smaller than when calculated from (11.35), and consequently P will be smaller when calculated from (11.35). Since the power input is smaller the value of k l a will be smaller for the given Δp.

The present problem will calculate the true value of v l and the ratio between the “true” P and P = P 0 calculated from (11.35). This will give us an analogue to N p,g/N p of Fig. 11.5.

  1. (a)

    For a fixed p t = 1.04 and a fixed Δp = 1.23 bar, v l = 199.2 m3 h−1 (the result for v l in Example 10.9 for t mix = 29 s, when the correction factor in (11.36) is not included). But n t = 10/199.2 = 0.0502, and when [n t, p t, Δp] = [0.0502, 1.04, 1.23] the correction factor is slightly smaller than 1, and v l < 199.2 m3 h−1.

    For different values of n t \( \subset \)(0,1) calculate the true v l by iteration of (11.36) – at most two iterations is enough to obtain convergence. Start with n t = 0.0502, but use also other values of v g at Δp = 1.23 bar to cover at least part of the interval 0 < n t < 1.

    Calculate P for the series of v g values used, and thereafter P/P(v l = 199.2 m3 h−1).

    How do these results influence k l a for increasing v g?

  2. (b)

    The results in (a) can be plotted in the same fashion as on Fig. 11.5.

    What is a reasonable variable to plot on the abscissa?

    Make your own conclusions concerning an analogue for N A in the RJH-mixer.

Rights and permissions

Reprints and permissions

Copyright information

© 2011 Springer Science+Business Media, LLC

About this chapter

Cite this chapter

Villadsen, J., Nielsen, J., Lidén, G. (2011). Scale-Up of Bioprocesses. In: Bioreaction Engineering Principles. Springer, Boston, MA. https://doi.org/10.1007/978-1-4419-9688-6_11

Download citation

Publish with us

Policies and ethics