2010_J Chem Tech Biotech_Bioprocess Scale-up

Embed Size (px)

Citation preview

  • 8/2/2019 2010_J Chem Tech Biotech_Bioprocess Scale-up

    1/15

    Review

    Received: 30 September 2009 Revised: 12 February 2010 Accepted: 12 February 2010 Published online in Wiley Interscience: 7 April 2010

    (www.interscience.wiley.com) DOI 10.1002/jctb.2387

    Bioprocess scale-up: quest for the parameters

    to be used as criterion to movefrom microreactors to lab-scaleMarco P. C. Marques,Joaquim M. S. Cabral and Pedro Fernandes

    Abstract

    Advancesin high-throughputprocessdevelopmentand optimizationinvolvethe rationaluse of miniaturizedstirredbioreactors,instrumented shaken flasks and microtiter plates. As expected, each one provides different levels of control and monitoring,requiring a compromise between data quantity and quality. Despite recent advances, traditional shaken flasks with nominalvolumes below 250 mL and microtiter plates are still widely used to assemble wide arrays of biotransformation/bioconversiondata, because of their simplicity and low cost. These tools are key assets for faster process development and optimization,

    provided data are representative. Nonetheless, the design, development and implementation of bioprocesses can presentvariations depending on intrinsic characteristics of the overall process. For each particular process, an adequate andcomprehensive approach has to be established, which includes pinpointing key issues required to ensure proper scale-up. Recently, focus has been given to engineering characterization of systems in terms of mass transfer and hydrodynamics(through gaining insight into parameters such as kLa and P/V at shaken and microreactor scale), due to the widespread useof small-scale reactors in the early developmental stages of bioconversion/biotransfomation processes. Within this review,engineering parameters used as criteria for scaling-up fermentation/bioconversion processes are discussed. Particular focus ison thefeasibilityof the application of such parameters to small-scale devices and concomitant usefor scale-up. Illustrativecasestudies are presented.c 2010 Society of Chemical Industry

    Keywords: scale-up; small-scale reactors; kLa; volumetric power consumption; fermentation; bioconversion

    NOTATIONa Specific interfacial area (m1)

    ai Initial specific surface area (m1)

    af Final specific surface area (m1)

    Bo Bond number, D2gWt1 ()

    d Maximum inside shaking flask diameter (m)

    D Well or vessel diameter (m)

    di Diameter of drops in size class i(m)

    Di Diffusivity (m2 s1)

    dmax Maximum drop diameter (m)

    dn Nozzle diameter (m)

    do Shaking diameter (m)

    d32 Sauter mean diameter (m)

    Fr Froude number, do(2N)

    2

    (2g)

    1

    ()g Gravitational constant (m s2)

    h Liquid height (m)

    k Number of size classes ()

    kL Mass transfer coefficient (m s1)

    kLa Volumetric oxygen mass transfer coefficient (s1)

    N Shaking frequency, stirring speed (s1)

    ni Number of drops ()

    Nf Pumping number ()

    NP Power number ()

    P Gassed power input (W)

    Po Ungassed power input (W)

    P/V Volumetric power consumption (W m3)

    Q Volumetric gas flow rate (m3 s1)

    Re Reynolds number, Ndo21, NT21 ()Sc Schmidt number, (Di)

    1 ()

    T Stirrer diameter (m)

    uo Nozzle velocity (m s1)

    V Filling volume (m3)

    vg Superficial gas velocity (m s1)

    Vo Flask volume (m3)

    vtip Impeller tip speed (m s1)

    W Width of turbine blades (m)

    We Weber number, N2T31 ()

    Wt Wetting tension (N m1)

    Subscripts

    c continuous

    T Tank

    Greek symbols

    Viscosity (Pa s)

    Correspondence to: Marco P. C. Marques, IBB-Institute for Biotechnology andBioengineering, Centre for Biological and Chemical Engineering, Instituto

    Superior T ecnico, Av. Rovisco Pais,1049-001 Lisboa, Portugal.E-mail: [email protected]

    IBB-Institute for Biotechnology and Bioengineering, Centre for Biological and

    Chemical Engineering, Instituto Superior T ecnico, Av. Rovisco Pais, 1049-001

    Lisboa, Portugal

    J Chem Technol Biotechnol2010; 85: 1184 1198 www.soci.org c 2010 Society of Chemical Industry

  • 8/2/2019 2010_J Chem Tech Biotech_Bioprocess Scale-up

    2/15

    Bioprocess scale-up: parameters to be used www.soci.org

    Density (kg m3)

    Local energy dissipation rate (W kg1)

    Volume fraction of the dispersed phase ()

    Interfacial tension (N m1)

    V Viscous dissipation term (m2 s2)

    INTRODUCTIONDespite sharing a common pattern, the design, development andimplementation of microbial processes present subtle variations,

    depending on the product, microbial strain, growth conditions,

    bioconversion/biotransformation conditions, among others. Thus,

    for a given product, process or facility, an adequate and

    comprehensive approach has to be established that encompasses

    the detailed characterization of the process and the timely

    identification of key process parameters likely to affect product

    yield, quality and consistency. These are to be kept as constant

    as possible throughout the scale-up process, in order to ensure

    success of the later task.1,2

    The development of microbial processes is strongly anchored

    in small-scale reactors, typically Erlenmeyertype flasks and bench-

    scale reactors, which are used for strain screening, media designand optimization and strategies of operation, ultimately aiming

    for the highest attainable productivity.3 In recent years, the

    range of small-scale vessels has increased to include multi-well

    plates (MWPs) with different levels of complexity and built-in

    devices, and miniature reactors that clearly emulate the larger

    vessels, but with a volume in the milliliter (or lower) range4,5

    Along with the technological developments that allowed for such

    hardware, efforts have been made to ensure the data gathered

    from experiments performed at these scales is reproducible

    throughout scales. Such efforts have relied on gaining further

    insight into mass (and heat) transfer and fluid dynamics at

    microliter scale, identification and validation of key parameters

    for scale-up, preferably from MWPs to bench-scale bioreactor,

    and predictive modelling.6,7,8 Knowledge gathered allows fulladvantage to be taken of the high level of parallelization

    provided by most miniaturized devices (in particular MWPs),

    and to speed up bioprocess development in a cost-effective

    manner.9,10,11 The task at hand is quite complex since there

    are several parameters influencing transport phenomena and

    chemical dynamics within a bioreactor. These parameters relate to

    mass transfer, mixing, partitioning, power input, shear induced by

    agitation, dilution rates, substrate and products concentration,

    nutrients, micronutrients and stabilizing agents, temperature

    and pH. Parameters of (bio)chemical nature are screened and

    optimized for bioconversion/growth kinetics. Physical parameters

    are, on the other hand, conditioned by process design and

    operational conditions. The parameters for scale-up are selectedwithin the range of physical parameters. Naturally favored

    candidates to fulfill such role are the process parameters

    or coefficients that are known to have some effect on the

    biological agent (enzyme or cell), particularly in their physiology.

    These include those affecting oxygen supply, heat transfer and

    mixing, namely aeration, agitation, mixing time, power input

    and oxygen mass transfer coefficient.2,12 In many cases, these

    physical parameters have to be combined with each other,

    or with other variables, in dimensionless numbers that are

    kept constant throughout scaling, therefore establishing scale-

    up criteria. In any case, the environmental conditions have to

    remain constant.1,2 Again, and despite the relevance of the

    scale-up issue in biotechnology, there is no straightforward

    and uniform strategy to tackle this matter. A suitable scale-up

    approach has therefore to be established again on a casuistic

    basis for a given product, process or facility.1,2,12 Most of these

    strategies, when bioconversion, fermentation or cell culture

    processes are involved, rely on the use of kLa or volumetric

    power consumption as criteria for scale-up, although constant Re,

    constant impeller tip speed and equal mixing and recirculation

    time are occasionally used.2,13 Given the relevance of the

    two former, they will be addressed in detail in this present

    work.

    SCALE-UP BASED ONKLAOxygen mass transfer coefficient

    Oxygen is a key substrate in most microbial processes of industrial

    relevance, where it can be required for growth, maintenance

    or production of metabolites.14,15 These processes are typically

    performed in an aqueous environment, but oxygen is sparingly

    soluble in water, roughly 0.272 mmol L1, at 25

    C and 101 kPa

    air pressure, and thus often becomes the limiting substrate.16 A

    suitable supply of oxygen to the liquid media, typically from air,

    is mandatory, but the process of mass transfer is influenced byseveral variables,such as physical properties of the fluids involved,

    operational conditions and geometry of the reactor. The oxygen

    transfer rate (OTR) can be increased by altering stirring speed and

    gas flow, which concomitantly alters the power input. Increasing

    the OTR is required to cope with the microbial oxygen demand,

    the oxygen uptake rate (OUR). It is possible to assess the rate

    limiting step of a microbial process, i.e. mass transfer or reaction

    limited, through the use of a modified Damkohler number (Da),

    which is calculated as the ratio between the maximum oxygen

    uptake and transfer rates.17

    Da =OURmax

    OTRmax(1)

    A large qO2 or low diffusivity leads to Da > 1, hence the

    process is mass transfer limited; oppositely, small qO2 or high

    diffusivity results in Da 1, thus the process is limited by the

    biochemical reaction rate.18 Oxygen transfer rate is thus a critical

    feature for the characterization of a given process and likewise

    an engineering parameter suitable for the design, selection and

    scale-up of bioreactors is to address OTR. One such parameter

    is the volumetric mass transfer coefficient kLa, which relates the

    oxygen mass transfer rate to the oxygen concentration gradient,

    according to (2). In a bioprocess the oxygen mass transfer rate can

    be described by12,16

    OTRC CL

    = kLa (2)where OTR is expressed as the molar flux of oxygen through the

    gasliquid interface; C is the dissolved oxygen concentration

    which would be in equilibrium with the gas phase, CL is the

    dissolved oxygen concentration in the bulk liquid, kL is the

    local mass transfer coefficient in the liquid phase, and a is the

    specific interfacial area. Although oxygen is transferred from the

    bulk gas phase to the bulk liquid phase, it is assumed that

    the gas phase resistance to mass transfer is negligible. Oxygen

    consumption through the process due to biochemical reactions

    can be considered using a biological enhancement factor, E,13,18

    J Chem Technol Biotechnol2010; 85: 11841198 c 2010 Society of Chemical Industry www.interscience.wiley.com/jctb

  • 8/2/2019 2010_J Chem Tech Biotech_Bioprocess Scale-up

    3/15

    www.soci.org MPC Marques, JMS Cabral, P Fernandes

    leading to an overall volumetric mass transfer coefficient KLa given

    by

    KLa = E.kLa (3)

    E incorporates the transport enhancement due to the oxygen

    uptake by microorganisms, alongside with opposing mass

    transfer resistances caused by layers of materials placed between

    the gas bubble and the bulk liquid phase, namely adsorbed

    surfactant and cells, and stagnant liquid film.18,19

    E has beenshown experimentally to change with the concentration of

    biomass, typically increasing with increased cell concentration,

    but with varying patterns, according to different strains and

    incubation media. In these dedicated experiments, E was within

    0.81.3, although values as high as 5 were also reported.13,18,19

    Nonetheless, and for most cases, the biochemical rate is not

    significantly larger than the mass transfer rate, hence it is usually

    assumed that E= 1. Details on this matter can be found in the

    recent review by Garcia-Ochoa and Gomez.13

    The volumetric mass transfer coefficient in bioreactors can be

    obtained experimentally or predicted using empirical correlations

    for kLa, or for kL and a, which are thoroughly described

    elsewhere,13 coupled to a model for the estimation of E. The

    introduction of the latter parameter in predictive determinationshas nevertheless been seldom reported, which may account for

    shifts between predicted and experimental data at high biomass

    concentrations. The use of constant kLa as scaling criterion is

    widely disseminated in conventional scales, from laboratory to

    production scales, encompassing volumes ranging from 1 L to

    1000 m3, and ultimately has also been found to be suitable down

    to the milli/micro-liter scale.68,2025

    Experimental determination of the volumetric mass transfercoefficient (kLa)

    Several methods have been developed to determine kLa

    experimentally.13,26,27 Within the scope of bioreactor design,

    several items are considered: stoichiometry, thermodynamics,microbial kinetics, transport phenomena (heat and mass transfer)

    and economics. While the first three items are scale-independent,

    transport phenomena and economics are extremely dependent

    on scale. The relevance of the transport phenomena in the design

    and scale-up of the bioreactor is particularly noticeable, since the

    overall rateof aerobic bioprocesses is generally controlled by mass

    transfer rates. The mass balance for the dissolved oxygen in the

    well-mixed liquid phase can be established as2,12,13,16

    dC

    dt= OTR OUR (4)

    where dC/dt is the oxygen accumulation rate in the liquid phase.

    The determination of OTR can be performed either when oxygenis being depleted by growing biomass (direct methods) or when

    no oxygen uptaketakesplace (indirectmethods).In thelatter case,

    equation (4) reduces to

    dC

    dt= OTR = kLa

    C C

    (5)

    The direct methods rely on oxygen probes, which allow for

    the determination of OTR through gas phase analysis or through

    the dynamic method. Until recently, only relatively bulky probes

    were available, which prevented oxygen monitoring in miniature

    vessels (MWPs, miniature/micro bioreactors), limiting its use to

    bench scale and larger bioreactors. Recently Erlenmeyer-type

    shaken flasks were adapted in order to be equipped with oxygen

    gas sensors, allowing for on-line determination of OTR in several

    parallel experiments in shaken vessels.28,29,30 Developments in

    fluorescence methods, (micro)fabrication techniques and optic

    fiber, have allowed for the implementation of sensitive dyes that

    can either be inserted into a patch and adhered inside a vessel,

    including individual wells from multiwall plates, yielding the so-

    called sensor spots, or incorporated onto the tip of fiber optic

    probes.3,4,11,25,31 37

    The indirect approach for the determination of OTR relies on

    chemical or physical methods. The most commonly used among

    the former is the cobalt-catalyzed oxidation of sulfite, which was

    optimized for application in miniature bioreactors (MWPs and

    shaken flasks),38 and is routinely used in such formats.33,39 A

    methodbasedonCO2 absorption is morerarely used.The physical

    methods, again relying on probes, allow for the dynamic method

    of OTR determination.13 Chemical methods, and in particular

    the sulfite method, may be biased due to modifications in fluid

    dynamics, fluid properties and surface tension, as a result of the

    addition of chemicals. They may therefore lead to misleading

    information and data gathered.13,27,40,41The use of fast enzymatic

    methods has also recently been introduced within the scope ofthe indirect approach. These are clearly designed for application

    in miniature systems, such as MWPs.42,43 These methods are

    based on the use of glucose oxidase and, preferably, of catechol

    2,3-dioxygenase. The latter method relies on a single-step well

    defined stoichiometric reaction, whereas the former requires

    calibration with the sulfite method (or a similar one) but on

    the other hand, all reagents are easily available. Among physical

    methods, the dynamic method is the most commonly used to

    evaluate kLa, because of its simplicity and relative accuracy. Both

    the absorption and desorption measurements give equal values

    of kLa under identical hydrodynamics conditions.44 When the

    characteristic time for the oxygen electrode and the characteristic

    time for the oxygen transfer process (1/kLa) are of the same

    magnitude, the dynamic response of the electrode has to beconsidered in the determination of kLa. A detailed description

    on the nature and limitations of the methods used for the

    experimental determination ofkLa is given by Garcia-Ochoa and

    Gomez.13

    Empirical correlations for the determination ofkLa in stirredtank reactors

    Both dimensional and dimensionless equations for the volumetric

    mass transfer coefficient as functions of different variables have

    been proposed.45 There are, however, considerable problems

    concerning the accuracy ofkLa estimation. Discrepancies between

    experimental data and those estimated from these equations are

    often found. The discrepancies are mainly found when kLa forreal broths are estimated from equations proposed for aqueous

    solutions. This can be due to the strong influence of the type and

    sizeof thebioreactor,the differentrangeof operational conditions,

    the system considered (solutions or real broths), or even the

    measuring method used.46,47 The addition of ions, hydrocarbons

    or temperature increase kLa, whereas the addition of surfactants

    or antifoams or increases in media viscosity decrease kLa, when

    compared with data obtained from water.26,4852

    Vant Riet26 proposed an overall correlation of kLa with

    volumetric power consumption and superficial gas velocity:

    kLa = C1 P

    VC2

    vgC3 (6)

    www.interscience.wiley.com/jctb c 2010 Society of Chemical Industry J Chem Technol Biotechnol2010; 85: 11841198

  • 8/2/2019 2010_J Chem Tech Biotech_Bioprocess Scale-up

    4/15

    Bioprocess scale-up: parameters to be used www.soci.org

    where P is the gassed power input, V is the liquid volume, vgthe gas superficial velocity and C1, C2 and C3 are constants that

    may vary considerably. Weuster-Botz etal.53 used correlation (6)

    to predict kLa in a magnetically driven stirred column, designed for

    parallel operation in an incubator chamber, operated as shaking

    flasks. Constants varied from 0.11 to 0.14, 0.06 to 0.37 and 0.73 to

    0.22, for C1, C2 and C3, respectively, according to differentturbines.

    Mass transfer coefficients of up to 0.34 0.05 s1 were obtained

    in the reactors, using defined medium for growing E. coli, using

    the dynamic gassing-out method.

    Montes etal.54 determined values of kLa in yeast broths

    (Trigonopsis variabilis) over wide ranges of both impeller speeds

    and superficial gas velocities, in three different mechanically-

    stirred, baffled reactors (2, 5 and 15 L). Experimental data were

    fitted using Equation (6) and the values for the parameters C2, C3and C1 were 0.35, 0.41 and 3.2 10

    3, respectively. Since most

    of the yeast broths behave as non-coalescent fluid, according to

    the authors, the correlation improved the prediction ofkLa values

    with respect to other generic correlations usually developed for

    strong coalescent and non-coalescent fluids. Additionally, Shin

    etal.55 verified that in high cell density cultures of fast-growing

    aerobes, such as recombinant E. coli, where the biomass mayincrease to more than 70 gL1, the oxygen availability can be the

    rate-limiting step of the fermentation process. Accordingly, the

    following correlation for kLa incorporating the effect of cell density

    (X) in oxygen transfer has been proposed:

    kLa = 0.0192

    P

    V

    0.55 vg0.64

    1+ 2.12X+ 0.2X20.25

    (7)

    Extensive details on this matter can be found elsewhere.13

    Another approach for the estimation of kLa relies on the use

    of empirical correlations incorporating dimensionless groups.13,45

    This approach has certain advantages because the correlations

    obtained for a known system can be used to estimate kLa in other

    systems with different dimensions.5658

    Although several correlationshave beendevelopedfor different

    systems, most of them are not specific to fermentation broths or,

    when developed with such a purpose, they do not take into

    consideration all the variations of parameters (surface tension,

    viscosity) throughout the time course of cultivation, which may

    hamper its effectiveness.

    Two-phase partitioning bioreactors have demonstrated signif-

    icant potential for enhancing the productivity of many biopro-

    cesses by overcoming issues of poor substrate solubility and

    toxicity. The oxygen mass transfer coefficient can also be evalu-

    ated in these systems. In order to take into account the effect of

    the organic phase and the organic phase volume fraction on kLa

    in aerated liquidliquid dispersions, empirical correlations havebeen proposed, assuming that the two liquid phases behave as a

    single homogeneous phase:59

    kLa =

    P

    V

    vg

    (1 ) (8)

    where is the volume fraction of the dispersed liquid phase, and

    , , and are numerical constants.

    Gomes etal.60 applied the correlation to the biotransformation

    of methyl ricinoleate into -decalactone by the yeast Yarrowia

    lipolytica. They showed that kLa had an influence on the

    aroma production; however, for the low hydrophobic substrate

    concentration used (1.08% v/v) and cellular density of 2 .0 107

    cells mL1, a minimal kLa value of 70 h1 was necessary to attain

    the maximum aroma production, 141 21 mg L1 (obtained at

    agitation andaeration rates of 400 rpmand 0.6 vvm, respectively).

    The numerical constants used were 650, 0.3, 0.7 and 1.7, for , ,

    and , respectively.

    Hydrodynamic studies in two-phase partitioning bioreactors

    have focused on gaining further insight into understanding the

    mechanismsrelatedwith formation of the interfacial areaavailable

    formass transfer, so that substrate supply (normally from the non-aqueous phase) does not become the rate limiting step of the

    process. The interfacial area has previously been correlated with

    the dispersed phase hold-up fraction and the Weber number.61

    The interfacial area available for mass transfer (a) is given by

    a =6

    d32(9)

    where d32 is the Sauter mean drop diameter. Accurate knowledge

    of the effectof bioreactor operating conditions on d32 is therefore

    very important. Knowledge ofd32 can also give an early indication

    of the stability of the liquidliquid dispersion created. The

    physicochemical properties of the media can influence both

    mean drop size and drop size distribution, as observed by Torres-Martinez, in the characterization of a multiphase system involving

    ionic liquids.62

    kLa determination in shaken devices

    One of the key challenges for shaken fermentation technology

    is to provide sufficient oxygen for the optimal growth of aerobic

    microorganisms.63 Adequate oxygen supply is crucial not only

    for industrial production, but also for meaningful screening and

    processdevelopment.64 Underoxygen-limitingconditionsaerobic

    microbes grow slowly, production of the intended metabolites is

    scarce, if any, and furthermore, unwanted synthesis of metabolites

    typical of anoxic conditions is prone to occur. Results obtained

    under such conditions are likely to be misleading, particularly forscale-up purposes.63 To study the effect of organism properties,

    medium composition or cultivation strategy on growth and

    production, incubation in a non-limiting oxygen environment

    is absolutely necessary. Otherwise wrong information about the

    variables under study might be obtained.65,66

    Several operational parameters affect the OTR in shaking

    devices, namely the shape and size of the vessel; the shaking

    frequency; the shaking amplitude; the shaking angle; the type

    of agitation (orbital or linear); and the filling volume. OTR is

    also influenced by the surface properties of the flask material,

    which may be either hydrophobic or hydrophilic; and by the

    physical chemical properties of the liquid (viscosity, oxygen

    solubility, diffusivity and surface tension, the latter being more

    noticeable in MWP).3,27,67

    Some particular setbacks are likely to occur when operating

    shaken vessels, specifically:

    1. The reproducibility of microbial growth might be poor.68

    2. When baffled vessels are used, small differences in depth

    and positioning of the baffles lead to significant differences

    in oxygen supply, growth and product formation of parallel

    cultivations.

    3. Out-of-phase phenomena might occur.67

    4. Shaking frequency in these flasks has to be reduced to avoid

    splashing of the liquid. Were droplets to reach the plug of

    the flasks, gas transfer limitations or contaminations might

    occur.63

    J Chem Technol Biotechnol2010; 85: 11841198 c 2010 Society of Chemical Industry www.interscience.wiley.com/jctb

  • 8/2/2019 2010_J Chem Tech Biotech_Bioprocess Scale-up

    5/15

    www.soci.org MPC Marques, JMS Cabral, P Fernandes

    Engineering features and correlations forkLa in shaken flasksand MWP

    Since bioprocess development (strain selection, strain enhance-

    ment,process optimization) is widely carried out in shakenvessels,

    and efforts towards engineering characterization of the oxygen

    transfer mechanisms in such devices have been undertaken.

    Maier and Buchs determined the maximum gasliquid mass

    transfer capacity in 250 mL shaking flasks on orbital shaking

    machines, using the sulfite oxidation method, by variation of

    the shaking frequency and diameter, filling volume and viscosity

    of the medium. The distribution of the liquid within the flask

    was modeled, taking as reference the intersection between the

    rotational hyperboloid of the liquid and the inner wall of the

    shaking flask.69

    The mass transfer within the shake flask is conditioned by

    two resistances: one due to the sterile closure of the vessel, the

    other due to the gasliquid interface. Owing to the resistance of

    the closure, the oxygen partial pressure within the flask is lower

    than the oxygen partial pressure of the surroundings. Mrotzek

    etal.70 concludedthat, under normal conditions, the mass transfer

    resistance of sterile closures is far smaller than the resistance of

    the gasliquid interface.The gas exchange through the sterile closure was described

    by the extended model of Henzler and Schedel,71 where besides

    Fick diffusion flow, other parameters and issues are taken into

    consideration, such as: (i) gas transfer by combined action of

    diffusion and convection due to non-equimolar mass exchange

    (Stefan flow); (ii) diffusion coefficients are notregardedas constant

    but are instead calculated as a function of the respective local gas

    concentration; and (iii) consideration of the water vapor flow.

    Operating with baffled flasks, and using suitable operating

    conditions, such as large shaken amplitudes, high OTRs can be

    obtained. In 250 mLshaking flasks witha filling volume of50 mL, a

    kLa of 400 h1 was obtained with a shaking amplitude of 2.5 cm. If

    theamplitude wasincreased to 5 cm,similar valueswere obtained

    at lower shaking frequencies. In these cases, the resistance of thesterile closure can become the limiting factor.

    Experimental data from given sets of experiments were fitted

    by least-squares after dimensional analysis,67 and the following

    proportionality for the maximum oxygen transfer capacity

    (OTRmax) was found

    OTRmax N0.84V0.84d0

    0.27d1.25 (10)

    Consequently, an increase of the maximum oxygen transfer

    capacity of a given system can be achieved by increasing the

    shaking frequency, reducing the filling volume, increasing the

    shaking diameter (d0) or reducing the maximum flask diameter (d)

    (this is only valid when the V1/3

    /dratio remains constant).Taking this into account, Maier etal.72 modeled the gas liquid

    mass transfer in shake flasks at water-like liquid viscosity, in flask

    sizes between 50 and 1000 mL. Relative filling volumes of 4 16%,

    shaking diameters of 1.25, 2.5, 5, 7, 10 cm andshaking frequencies

    of 50 500 rpm were tested. Furthermore, the previous model of

    the gasliquid mass transferEquation (10)was extended to a two

    sub-reactor model, to account for different mechanisms of mass

    transfer in the liquid film on the flask wall, and the bulk of the

    liquid rotating within the flask. The two-reactor system approach

    consists of a stirred tank reactor (bulk liquid) and a film reactor

    (film on flask wall and base). The mass transfer into the film on the

    flask wall and base at in-phase operating conditions, is described

    by Higbie penetration theory. Two differentmass transfer theories

    were applied to successfully describe the mass transfer into the

    bulk liquid: a model by Kawase and Moo-Young73 and a model by

    Gnielinski.74 Extensive details can be found elsewhere.72

    The agreement between the new modeling approach and the

    experimental data was within 30%. The applicability of the

    latter models to a biological system was shown using a Pichia

    pastoris culture. The OTRmax was determined using the sodium

    sulfite method which displayed a correlation factor of 2.8, when

    compared with the OTR determined using an oxygen limited

    P. pastoris culture, under variation of the operating conditions

    (250 mL shake flask with filling volumes of 15, 25 and 40 mL,

    under shaking frequencies of 50500 rpm with 5 cm of shaking

    diameter).72 Moreover, the volumetric mass transfer coefficient

    models allowed a significant agreement between experimental

    dataandmodelpredictions,asreflectedbyacorrelationcoefficient

    of 0.88. The maximum volumetric mass transfer coefficient of the

    experimental investigation was found to be 0.157 s1 inthe 50 mL

    flask, at a relative filling volume of 4%, a shaking frequency of

    450 rpm and a shaking diameter of 7 cm.

    Liu etal. correlated the experimental data for the determination

    of OTR with a multivariable power correlation in carotenoid

    (astaxanthin) production by the red yeast Phaffia rhodozyma.15

    The constant parameters were derived by linear regression of the

    data, resulting in the following equation:

    kLa = 0.141N0.88

    V

    V0

    0.80(11)

    where Vo is the flask volume. Results showed a direct linear

    correlation between carotenoid yield and OTR (OTR from 0 to

    690 mg L1 h1), indicating that carotenoid production is limited

    by oxygen transfer. Mantzouridou etal.75 also investigated the

    effect of oxygen transfer rate on -carotene production by

    Blakelsea trispora in shake flasks. The results indicated that the

    concentration of-carotene (704.1 mg L1) was highest in culture

    grown at maximum OTR of 20.5 mmol L1 h1. Moreover, OTRlevelshigher than20.5 mmol.L1h1 werefound to be detrimental

    to cell growth and pigment formation.

    Besides increasing the diameter or frequency of shaking or

    decreasing the filling volume, one possibility to achieve high

    maximum oxygen transfer capacities (OTRmax) is to modify the

    usual round geometry of the shaken vessel. As shown by many

    groups, changing the geometry of the vessel (i.e. from round-

    bottomed base to square-bottomed base) and/or introducing

    baffles into shake flasks resultin a significant increase in maximum

    oxygen transfer capacity.76 The OTRmax can be increased up to

    5 10-fold even at low shaking frequencies.

    Alternatives for the well bottom design have been established,

    among them the square shaped well. This type of well geometryhas been investigated by Duetz and Witholt42,77 and Duetz etal.78

    The effects found for square wells are comparable with those in

    baffledshake flasks.Eventhough squarewells allowfor an increase

    in OTRmax of roughly 100% when compared with round wells, the

    aforementioned problems described for shake flasks (splashing

    and out-of phase phenomena) may limit the utilization of square

    deep well plates as cultivation vessels.

    More recently, Funke etal.79 presented well configurations with

    a geometry thataimed at avoiding splashing while simultaneously

    allowing for a high filling volume. Different plate formats were

    studied, where variation were achieved by increasing the number

    of edges and rounding the edges in square geometry plates,

    rounding edges in pentagonal shape wells, star and flower

    www.interscience.wiley.com/jctb c 2010 Society of Chemical Industry J Chem Technol Biotechnol2010; 85: 11841198

  • 8/2/2019 2010_J Chem Tech Biotech_Bioprocess Scale-up

    6/15

    Bioprocess scale-up: parameters to be used www.soci.org

    shape wells, among others. Dissolved oxygen tension (DOT) was

    measured in an experimental setup described by Samorski etal.80

    The well geometry proved influential for OTRmax. By introducing

    such baffles, an OTRmax in excess of 100 mmol L1 h1 (kLa over

    600 h1) was obtained, roughly double that of round 48-well

    MWPs. In conventional MWPs, values for OTRmax reported so far

    have not generally exceeded 50 mmol L1 h1.27 Values of the

    OTR in stirred tank reactors can reach 500 mmol L1 h1, but

    overall, in standard batch fermentations, the OTR does notexceed

    100 mmol L1 h1.26 The final well design, a six-petaled format,

    provided the best compromise among different levels of baffling,

    allowing simultaneously for a stable liquid height at the well

    center during shaking (strong), high OTRmax independently of

    filling volume and shaking frequency (moderate), and high filling

    volume without splashing.79

    Nonetheless, there are two common problems associated with

    the use of microtiter plates for carrying out fermentations. These

    are evaporation and cross-contamination due to spillover.In order

    to avoid these drawbacks, MWPs can be sealed with adhesive

    tape. The use of sealing tape leads to airtight closure of the wells

    decreasing oxygen transferto the reactor welland consequentially

    to the reaction medium.11

    Doig etal.81 modeled the volumetric oxygen transfer coefficient

    in MWPs using dimensionless groups. The basis for the modeling

    was the experimental measurements of airliquid specific surface

    area, determined both by the rate of evaporation and by high-

    speed video photography.

    Flowbehaviorwas modeled separatelyaccording to the specific

    microplate configuration (24, 96 and 384-well), and the following

    correlations, respectively, were obtained:

    af

    ai= 2.895 Fr0.86Bo0.03 (12)

    af

    ai= 1.092 Fr0.64Bo0.15 (13)

    af

    ai= 0.607 Fr0.51Bo0.18 (14)

    The different models for predicting specific air liquid surface

    area (af/ai) converged, with most data points lying within 25%

    of the predicted values.

    The kLa values ranged from about 0.005 to 0.055 s1 (shaking

    amplitude of 3 mm and shaking frequencies ranging from 200

    to 800 rpm) and are within the range reported by other authors

    for these devices.5,35,78,82 Since the specific surface air liquid area

    increased by a maximum factor of 4, and kLa increased by up to

    10, it is clear that the liquid side oxygen transfer coefficient kL was

    also affected by shaking conditions. Hermann etal.82 observed

    an increase in kL from about 5.5 10

    5

    to 1.4 10

    4

    ms

    1

    asshaking frequency was increased from 200 to 800 rpm at 25 mm

    amplitude. Values calculated by Doig etal.81 for kL varied from

    2.8 105 ms1 to 8.3105 ms1 (shaking amplitudes of 3 mm

    to 8 mm, in 24, 96 and 384 well microtiter plates).

    Correlations for kL typically encompass Reynolds and Schmidt

    numbers,and areusuallyin theformSh = a RebScc,where b ranges

    from 0.7 to 0.8 and c is usually 0.33.81 The optimized correlation

    for all three microplate geometries is

    Sh = 0.19 Re0.68 Sc0.36 (15)

    All the experimental data were modeled within 30% de-

    viation, using correlation (15). Combining equation (15) with

    Equations (12)(14), an overall correlation for predicting the vol-

    umetric oxygen transfer coefficient kLa, in round bottom plates

    becomes

    kLa = 31.35DiaiRe0.68 Sc0.36 FrxBoy (16)

    where Di is the diffusion coefficient, ai is the initial specific surface

    area, and x and y are constants depending on the microplate

    geometry (according to Equations (12)(14).

    Islam etal.83

    modified Equation (16) in order to predict kLa insquare well microtiter plates. The correlation obtained was:

    kLa = 3.94 104

    T

    D

    ai Re

    1.91eaFrb

    (17)

    where the values for a and b were 1.66 and2.47 for48 rectangular,

    flat wells; 0.70 and 1.51 for 24 square wells, round bases; and 0.88

    and 1.24 for square wells, pyramidal bases. The predicted values

    were in good agreement with the experimental data for kLa.

    Case studies: scale-up from shaken devices and miniaturestirred reactors based inkLa similarity

    In previous studies84 kLa was identified as the key engineering

    parameter for characterization of an E. coli based process forheterologous protein expression, in microtiter plate format. A

    scale-up to 7.5 and 75 L stirred tanks was performed, using as

    criterion kLa fixed at either 0.069 s1 or at 0.015 s1. At 0.069 s1

    boththe fermentation profile (biomass concentrationand glycerol

    consumption) and product yield were identical in all scales,

    enabling a 15 000-fold quantitative scale-up, representative of

    fermentation performance. At 0.015 s1, due to poor gasliquid

    distributions observed within the larger stirred tanks at matched

    kLa, the overall fermentation profile was not reproduced.

    Micheletti etal.6 studied the scale-up of aerobic fermentation

    of E. coli JM107:pQR706 for overexpressing transketolase (TK),

    from microtiter plates to stirred reactors. Using the correlation

    displayed in Equation (17), a kLa of 0.079 s1 at 1000 rpm for

    a 96 deep well microtiter plate was obtained. Maintaining the

    volumetric oxygen mass transfer coefficient, it was possible to

    scale-up the process directly from the microwell reactor to a lab-

    scale reactor (1.4 L). Apart from some differences in duration of the

    lag phase observed when the two scales were compared, similar

    values ofmax and final biomass concentration were obtained.

    Likewise, the rates of L-erythrulose formation when the cells

    from the respective fermentations were used for the subsequent

    HPA(-hydroxypyruvate) and GA (glycolaldehyde) bioconversion,

    were also similar.

    Zhang etal.8 used computational fluid dynamics (CFD) to

    provide a detailed characterization of fluid mixing, energy

    dissipation rate and mass transfer in single well bioreactors,

    from deep square 24-well and 96-well microtiter plates. The CFDsimulations showedthat liquidmixing is more intensive in 96-well

    than in 24-well bioreactors, due to the vertical movement of the

    bulk fluid, in addition to the rotational movement. Liquid motion

    was strongly dependent on the orbital shaking amplitude which

    generally has a greater impact than the shaking frequency.

    Predicted kLa values were compared with experimental results

    obtained from dynamic gassing out experiments using a fibre-

    optic dissolved-oxygen probe. The resulting kLa values measured

    were 0.036 s1 and 0.023 s1, at 1000 rpm and 500 rpm, respec-

    tively, in 96-well MWPs, at an orbital shaking amplitude of 3 mm.

    The corresponding predictedvalues, 0.065 s1 and 0.056 s1 ,were

    higher. The discrepancy was ascribed to some features of the dy-

    namicmethod,namelytotheprobesensitivitytowardstheshaking

    J Chem Technol Biotechnol2010; 85: 11841198 c 2010 Society of Chemical Industry www.interscience.wiley.com/jctb

  • 8/2/2019 2010_J Chem Tech Biotech_Bioprocess Scale-up

    7/15

    www.soci.org MPC Marques, JMS Cabral, P Fernandes

    movement of the microtiter plates, as well as to the positioning of

    the probe andto the mixingtime. CFDstudies could give more ac-

    curate values if mediumlosses due to evaporation were taken into

    account, and if the actual physical properties of the fermentation

    media (viscosity, density) were used instead of water, as input for

    performing the calculation. Batch cultures ofE. coliDH5 showed

    similar maximum specific growth rates and final biomass yields

    in shaken 24-well MTP and Erlenmeyer bioreactors, and in stirred

    miniature and 20 L bioreactors at matched kLa values.

    The biotransformation of benzaldehyde to l-phenyl acetyl

    carbinol was scaled-up from a 100 mL shake flask to a 5 L reactor,

    using resting cells of Saccharomyces cerevisiae.85 The following

    correlations were used for the prediction of kLa in the growth

    medium and in the biotransformation medium, respectively,

    kLa = 0.024

    P

    V

    0.725v0.892g (18)

    kLa = 6.99 106

    P

    V

    1.14v0.365g (19)

    The volumetric oxygen mass transfer coefficient varied from0.010.07 s1 in the growth medium to 0.0050.02 s1 in the

    biotransformation medium. Maintaining the kLa constant as

    scale-up criteria, both for the cell growth process and for the

    biotransformation, a 50-fold scale-up was achieved.

    Gill etal.7,86 studied the influence of kLa on the fermentation

    of E. coli TOP10 pQR239 in a 100 mL reactor. A correlation was

    developed for the miniature reactor

    kLa = 0.224

    P

    V

    0.35v0.52g (20)

    which was validated using the dynamic gassing out technique.

    For scale-up to a 2 L bioreactor, different values of kLa were

    used, varying from 0.06 to 0.11 s1. The results showed that therewas good agreement between cell growth and dissolved oxygen

    tension profiles across the range ofkLa values studied, compared

    with experiments at matched P/V values. The trend in oxygen

    depletion during the growth phase and the time taken for both

    systems to become oxygen limited were similar and reproducible

    in both cases, thus resulting in a satisfactory 20-fold scale-up.

    Using an E. coli JM107:pQR706 overexpressing transketolase

    system, Micheletti etal.6 showed that kLa is the most suitable

    scale-up parameter from a 24-well microtiter plate (1 mL filling

    volume) to a 2L stirred tank reactor. On the other hand, the

    prediction for the power consumption in the mechanically stirred

    bioreactor (3.64 W m3) working under operational conditions

    that allowed for growth patterns that matched those in shakenvessels was roughly 10-fold lower.

    More recently, Marques etal.5 achieved an 8000-fold scale-up

    for the side-chain cleavage of -sitosterol performed by whole

    cells of Mycobacterium sp. NRRL B-3805 in 24-well microtiter

    plates to a 5 L reactor. Scale-up was performed based on kLa

    similarity at 0.058 s1 (at a shaking frequency of 250 rpm and

    filling volume of 0.5 mL). Similar profiles (glycerol consumption

    and 4-androstene-3,17-dione (AD) production) were achieved in

    both scales validating multi-well plates as a small-scale reactor

    for performing complex bioconversions. Nonetheless, there was

    a consistent gap between values obtained in the multi-well

    plate system and the bench-scale reactor for AD production and

    biomass formation visible also in the levels of oxygen depletion.

    The behavior was ascribed to diverse intrinsic hydrodynamic

    conditions in the different reactor configurations, since mass

    transfer plays an important role, both for oxygen and substrate

    uptake.

    Both MWP as well as miniature bioreactors proved effec-

    tive as starting platforms for scaling up to bench/pilot scale

    biotransformation/fermentation, but further studies in different

    environments would be welcome that would hopefully further

    validate this approach. Case studies are relatively scarce, where

    miniature bioreactors are concerned. This can be partly ascribed

    to the wider dissemination and lower cost of MWP platforms,

    when compared with miniature bioreactors, whose availability in

    the market is clearly lower.11 The studies referred to in this paper

    related to the use of miniature bioreactors are based on in-house

    developed prototypes. The use of the simpler MWPplatforms may

    be of limited use when pH shifts take place during the process,

    given the lack of mechanisms for pH control, only available in

    more complex platforms.11 Processes involving significant shifts

    in broth rheology (i.e. production of polymers) or viscous envi-

    ronments may be more adequately dealt with using miniature

    bioreactors, since these may provide more efficient mixing.

    SCALE-UP BASED ON VOLUMETRIC POWERCONSUMPTIONPower consumption is a key parameter in (bio)chemical engineer-

    ing,i.e. anengineering characterization parameter. Concomitantly,

    it is a strong candidate for use as a criterion for (bio)reactor de-

    sign and process scale-up. Often referred to as volumetric power

    consumption (P/V), it is defined as the amount of energy required

    to generate movement of a fluid within a vessel in a given pe-

    riod of time. Along with the power actually drawn by the fluid,

    relevant to the outcome of a given process, further power is re-

    quired to account for energy losses, mostly due to friction, and

    power consumption by motorsand gearboxes. This excess power,

    although often relevant in terms of overall power consumption,

    is not considered for design or scale-up of the process. The vol-

    umetric power consumption is representative of the turbulence

    degree and media circulation in vessels, and influences heat and

    mass transfer, mixing and circulation times.87 Constant volumetric

    power input wasapplied successfully as scale-up parameterfor the

    early industrial penicillin fermentations (1 hp gallon1, equivalent

    to1.8 kW m3), andin fermentations with lowenergy inputs,88 but

    it is limited in fermentations requiring high energy inputs, such

    as recombinant E. coli cultures,1 possibly due to high associated

    costs and to high shear stress in the larger-scale stirred vessels.

    AccordingtoRushton etal.,89,90 thepowerinputfortheagitation

    of a non-aerated mixture, Po is characterized by the dimensionless

    variable power number (NP):

    P0 = Np N3T5 (21)

    where N is the stirrer speed and T the stirrer diameter. The

    power number depends on other dimensionless groups such as

    Reynolds number and Froude number, as well as on the number

    of agitator turbines. Thepowerconsumption in ungassed systems

    is always higher than the power consumption in gassed systems,

    since aeration significantly influences the power drawn from the

    impeller by the fluid.7 The effect of aeration has been studied

    extensively by Nienow etal.,91 Oosterhuis and Kossen,92 Yawalkar

    etal.93 and Gogate etal.46 It has been shown that the gassed

    power input is usually 3040% of the ungassed power input,

    www.interscience.wiley.com/jctb c 2010 Society of Chemical Industry J Chem Technol Biotechnol2010; 85: 11841198

  • 8/2/2019 2010_J Chem Tech Biotech_Bioprocess Scale-up

    8/15

    Bioprocess scale-up: parameters to be used www.soci.org

    depending on the type of impeller and aeration rates used.79

    Cavities (gas pockets) are formed behind the impeller blades in

    gassed systems, resulting in different densities of the fluid under

    gassed and ungassed conditions.94,95 Too high gas flow is to be

    avoided since this may lead to cavities extending between blades,

    which result in flooding and mechanical instability. Cavities may

    be minimized,hence gas handling capacityimproved, withthe use

    of adequate impellers, such as the Rushton turbine. Alternative

    blade impeller configurations (i.e. concave blades) have also been

    introduced.96,97

    Hughmark98 suggests the following relationship for estimating

    the volumetric power consumption:

    P

    Po= 0.1

    N2T4

    gWV23

    1

    6 Q

    NV

    14

    (22)

    where W is the width of turbine blades and Q is the volumetric

    gas flow rate. Over the years several modification were proposed

    in order to improve this model. Such modifications included

    taking into account the number and type of impellers and reactor

    dimensions, among others.99,100,101

    Along with the use of predictive correlations, power con-

    sumption may be determined experimentally through the use

    of electrical or calorimetric measurements, or through the use of

    dynamometers, torquemeters and strain gauges.87

    In bench- and pilot-scale bioreactors typically used for the

    fermentation of bacteria and fungal micro-organisms, the power

    consumption ranges from 1 to 3 kW m3.102,103

    Shaken flasks

    Although well established for bench-scale and above reactors,

    only recently have dedicated efforts been made to provide a

    suitable characterization of the volumetric power consumption in

    miniaturized devices, and to take advantage of this parameter forthe engineering characterization of said devices. Most reports on

    the experimental determination of power consumption in shaking

    systems are by the group of Jochen Buchs at RWTH Aachen. Buchs

    etal.104,105 measured the power consumption in shaken flasks at

    high and low-medium viscosities. The experimental assembly was

    a simplerotary shaking machine fixed to a frame,combinedwith a

    torque sensorattached tothe powering drive.Torque andshaking

    speed were monitored and correlated by the following relation:

    P

    V= Ne

    N3d4

    V23

    = C3N3d4

    V23

    Re0.2 (23)

    where Ne

    is the modified Newton number for shake flasks.The correlation implies that the specific power consumption is

    dependent on the shaking frequency according to P N2.8. Such

    a correlation is typical of that found in unbaffled agitated tank

    reactors. The model includes C3 as the only fitting parameter

    using least-squares non-linear fitting for the description of all the

    experimental results obtained in their study, having a value of

    1.94.

    The values for volumetricpower consumptioncalculatedby this

    empirical correlation and the experimentally measured values fit

    within a deviation range that does not exceed 30%. In the range

    0.01 to 0.2 kW m3, with shaking frequency from 80 to 380 min1,

    filling volumes of 4% to 20% of nominal flask volume and

    shaking diameter from 2.5 to 5 cm, larger discrepancies are found

    between calculated and measured values. The corresponding

    data was gathered at low shaking frequency (80 to 120 min1)

    and, therefore, at low Reynolds numbers (Re 500 to 5000),

    possibly within the transition from laminar to turbulent flow, a

    feature that could account for the increased deviation between

    predicted and experimental values.

    The discrepancy between experimental and predicted data at

    low power consumption was overcome in liquids with viscosities

    between 0.8 and200 mPa s,105 where a correlation was found that

    took into consideration all the flow regimes:

    Ne = 70Re1 + 25Re0.6 + 1.5Re0.2 (24)

    This correlationconsists of a laminar (Re1), a transition (Re0.6),

    and a turbulent term (Re0.2). From the power number variation

    with the flask Reynolds number, two flow conditions were

    identified: in-phase conditions, where the bulk of the liquid

    in the flask circulates in phase with the shaking platform; and

    out-of-phase conditions, where only a minor fraction of the

    liquid is actually moving along the flask wall.67 The out-of-phase

    conditions lead to a decrease in volumetric power consumption,

    mixing gas/liquid mass transfer. In order to systematically describethe in-phase and out-of-phase conditions, a new non-dimensional

    number, called the phase number (Ph), was derived based on an

    analogy to a partially filled, rotating horizontal drum, and can be

    expressed as

    Ph =d0

    d

    1+ 3log10

    (2N)

    d2

    4

    1

    14

    V0.33

    d

    2

    2

    (25)

    Buchs etal.98 established that all operating conditions where

    Ph > 1.26arein-phasewhileout-of-phaseconditionsareobserved

    for Ph < 1.26, which are prone to occur when large flasks or high

    viscosityfluids are used.11 Nonetheless, out-of-phase conditions in

    largeflasks at water-likeviscosities arepossibleif theflask is quickly

    accelerated.106 This may occur, if (i) the shaking machine has a

    strong engine, (ii) the load of the machine is low (small number of

    flasks placed on the shaker table or (iii) the shaking flasks which

    wereremovedfromtheshakingtable(e.g.forsampling)areplaced

    back on the running machine.

    This phenomenon of out-of-phase operating conditionsmay be

    of great practical relevance, since working under such conditions

    will significantly reduce the oxygen transfer and mixing intensity,having a strong impact on strain and medium development. The

    flow in a shaking flask tends to be in-phase at large shaking

    diameter, low viscosity, large filling volume, higher shaking

    frequency and small number and size of baffles.

    Kato etal.107 used a calorimetric method, based on the method

    developed by Sumino etal.,108 for assessing power consumption

    in larger shaken flasks (nominal volume up to 20 L). Despite the

    increase in volume, the order of magnitude ofP/V is the same as

    for smaller scale shaking flasks, up to a nominal volume of 2 L and

    having different geometries.104

    Similar studies have been conducted reporting a stronger

    dependency between power consumption and shaking speed,

    P N.5.75 Nonetheless, a narrower range of shaking frequencies,

    J Chem Technol Biotechnol2010; 85: 11841198 c 2010 Society of Chemical Industry www.interscience.wiley.com/jctb

  • 8/2/2019 2010_J Chem Tech Biotech_Bioprocess Scale-up

    9/15

    www.soci.org MPC Marques, JMS Cabral, P Fernandes

    between 100 and 200 rpm was used, compared with the

    100 300 rpm range of Kato etal.1 This may also account for

    the volumetric power consumption reported in either work, from

    less than 2 kW m3 to 5 kW m3, respectively. Therefore, it can be

    concluded that the increase in power consumption is much more

    pronounced in the region of low shaking frequencies, between

    100 and 200 rpm.

    Raval etal.109 compared the torque method and temperature

    methodforthedeterminationofpowerconsumptionindisposable

    shaken bioreactors of 2 L polycarbonate (PC) bottles and 20 L

    polypropylene (PP). Data points collected by both methods were

    within30%tolerance range.Values of3.5 kW m3 wereobtained

    in 2 L flasks with filling volume of 250 mL at 300 rpm (shaking

    diameter 50 mm).

    Another approach for estimation of the volumetric power

    consumption was proposed by Zhang etal.103 Using CFD

    techniques,theseauthorswereabletopredictpowerconsumption

    when 250 mL shake flasks were operated between 100 and

    300 rpm, with shaking diameters between 20 and 60 mm and

    filling volumes between 25 and 100 mL. CFD models suggest that

    P/V can be correlated with the filling volume or with the shaking

    frequency according to P/V V0.7

    and P/V N2.7

    , respectively.Both correlationsare in agreementwith those suggestedby Buchs

    etal.104 Nonetheless, for lower shaking diameters the exponent

    in the shaking frequency correlation decreased to 1.6, suggesting

    that the applicability of these correlationsare limited to a range of

    amplitudes. Values predicted were between 40 and 1200 W m3.

    The powerconsumptionestimates werefoundto be moresensitive

    to changes in shaking amplitude than to frequency

    Multi-well plates

    Despite the widespread application of MWPs, currently there is

    no experimental report on power consumption in these devices,

    mainlyduetolackofcommercialtorquemetersorothermeasuring

    methods withsufficient sensitivity to perform suchmeasurements.In order to overcome this situation, Zhang etal.8 proposed the use

    of CFD to obtain estimates of power consumption. These authors

    applied the methodology followed by Zhang etal.103 where the

    local energy dissipation rate was obtained using

    =v

    (26)

    where V is the viscous dissipation term. The power consumption

    based on the fluid friction is calculated from

    P

    V

    =

    V0

    vdV

    V

    (27)

    where V, can be expressed in terms of shear rates:103

    V = 2

    u

    x

    2+

    v

    y

    2+

    w

    z

    2+

    u

    y+

    v

    x

    2

    +

    v

    z+

    w

    y

    2+

    u

    z+

    w

    x

    2(28)

    where U(u,v,w) is the velocity vector(velocity component) andx(x,

    y, z) is the moving grid spatial vector (spatial component).

    The results showed that power consumption and energy

    dissipation rates in the shaken microwells are strongly affected

    by the size of the well and the filling volume, as well as by the

    shaking frequency and amplitude. These effects are higher in 96-

    well microtiter plates than in 24-well microtiter plates under the

    same operating conditions.

    Moreover, two situations can be distinguished:

    1. The power consumption in the 24-well reactor does not

    increase linearly as theshaking frequency is increasedat 3 mm

    shaking diameter. Up to roughly 800 rpm, there is a decreasein power consumption, presumably due to out-of-phase

    conditions. The estimated volumetric power consumption

    is more sensitive to changes in shaking amplitude than to

    frequency.

    2. The volumetric power consumption at orbital shaking ampli-

    tudes of both 3 mmand 6 mmincreased linearly with increase

    in shaking frequency. Increasing the shaking amplitude from

    3 mm to 6 mm, led to an increase in the volumetric power

    consumption of 1000 2000 W m3 using shaking frequencies

    in the range of 800 1000 rpm.

    The volumetric power consumption was compared for a 24-well

    microtiter plate, 96-well microtiter plate, 20 L stirred reactor, 6 mL

    miniature stirred reactor110 and a 250 mL shake flask102 at feasibleoperating conditions.

    The miniature stirred reactor closely matched the configuration

    of conventional stirred reactors, only scaled down to 6 mL

    scale.11 The P/V was similar in all reactors with the exception

    of shaken flasks and the 24-well microtiter plates. The values used

    for microtiter plates were obtained by CFD studies in specific

    conditions; nonetheless similar values of P/V can be reached in

    shaken systems.

    Case studies: scale-up from shaken devices and miniaturestirredreactorsbasedin specificpowerconsumptionsimilarity

    Aqueous systems

    Despite the large number of studies on scale-up of fermentationprocesses andsome on bioconversion processes, examples on the

    use of volumetric power consumption as key parameterare scarce.

    Gill etal.7 developed a prototype of a microreactor that enables

    parallel operation of 4 16 independently controlled experiments.

    Each microreactorhas a maximum working volumeof 100 mL and

    is equipped with a magnetically driven six-blade Rushton turbine,

    with a stirring range of 100 2000 rpm. Growth ofE. coli TOP10

    pQR239was scaled-upfromthismicroreactor toa 2 L reactor fitted

    with two six-blade Rushton turbine impellers111 using as constant

    the power consumption. P/V values used in this study were 657,

    1487 and 2960 W m3.

    At the lowest P/V value of 657 W m3, the 2 L reactor signif-

    icantly underperforms compared with the miniature bioreactor,achieving a final biomass concentration of almost 3 g L1 less

    (Xfinal in the miniature bioreactor was 5.6 g L1). This is likely to be

    the result of operating at a reduced agitation rate, and the poor

    gasliquid dispersion that was observed at the 2 L scale under

    these operating conditions. Given the poor oxygen transfer at the

    2 L scale, the dissolved oxygen tension reached zero much earlier,

    and cell growth was clearly seen to become oxygen limited. The

    performance of the 2 L reactor was improved at higher P/Vvalues

    (>1000 W m3), with very similar max and final biomass con-

    centration values obtained at both scales. The trends for oxygen

    depletion during the exponential growth phase and for the time

    taken forbothsystems tobecome oxygenlimited were similar and

    reproducible at the twohigher P/Vvalues(1487and 2960 W m3).

    www.interscience.wiley.com/jctb c 2010 Society of Chemical Industry J Chem Technol Biotechnol2010; 85: 11841198

  • 8/2/2019 2010_J Chem Tech Biotech_Bioprocess Scale-up

    10/15

    Bioprocess scale-up: parameters to be used www.soci.org

    The volumetric power consumption was also used to scale-up

    the alginate production using Azotobacter vinelandii cells, from

    500 mL shake flasks to 14 L reactor.112 The study showed that

    when an initial power draw of 0.27 kW m3 was applied, a specific

    growth rate of 0.16 h1 was obtained in a stirred reactor, which

    roughly doubled the specific growth rate obtained in shake flasks

    (0.09 h1). Moreover, differences in the broth viscosity, concentra-

    tion profiles and molecular weight of the alginate were observed.

    In order to overcome these, the initial P/V in the reactor or along

    the time course of the cultivation was reduced, which ultimately

    allowed oneto match the molecularcharacteristics of the alginate

    obtained in shake flasks. This approach further highlights the

    potential limitations of shaken vessels as starting platforms for

    scaling-up when significant modifications in medium rheology

    take place throughout the process. This study was based on the

    theoretical analysis of power consumption in shake flasks. The

    power consumption in shake flasks was estimated from extrapo-

    lationof data reported by Buchs etal.104,105This is a key feature, as

    thechangesin power inputaffect theOTR,which in turn affects the

    molecular characteristics of alginate. Further work was performed

    in order to gain insight into the power consumption along cell

    growth for the same alginate-producing system. Pena etal.113

    showed that the power consumption increased exponentially

    during fermentation, achieving a maximum value of 1.4 kW m3

    after 40 h cultivation. This increase was due to the increased

    viscosity of the culture broth, which resulted from the increase

    in the molecular mass of alginate and polymer concentration. In

    this period, a maximal alginate concentration of 5 kg m3, with a

    maximum molecular weight of 550 kDa was obtained.

    Although the viscosity increasedin theperiod from 40 to 70 h, a

    slight drop in the power consumption was observed, leading to a

    value of 1.2 kW m3 at 70 h cultivation. This behavior waspossibly

    due to out-of-phase conditions, which could also account for the

    results obtained previously by Reyes etal.112

    Pena etal.114 tried toreproduce themean molecularmassof the

    alginates obtained in shake flasks, in a stirred reactor maintainingP/V constant. A 500 mL Erlenmeyer flask was used, containing

    1/5 filling volume. The power consumption during alginate pro-

    duction increased exponentially from 0.18 to 1.4 kW m3 during

    the first 40 h of culture and remained practically constant during

    the rest of the fermentation. The exponential profile of the power

    consumption(from0.2to1.2 kW m3)wassimulatedalongthefer-

    mentationina14 Lbenchreactor,containing10 LofBurkmedium,

    by adequately controlling the agitation rate from 250 to 515 rpm.

    Further dissemination of the evaluation of the validity of

    constant P/V as criterion for scaling-up is dependent on the

    availability of methodologies and tools for easily assessing this

    parameter in miniaturized devices.

    Studies in two-phasesystems

    All the above examples are for aqueous systems. When substrates

    that have low water solubility are involved, two-liquid phase

    systems can be used. In these cases, P/V is particularly important

    as drop diameter (and hence the interfacial area available for

    mass transfer between phases) depends on the maximum energy

    dissipation rate or volumetric power consumption. Moreover, for

    efficient extraction it determines also the mean drop size and the

    dispersion of the immiscible organic phase.115

    In a stirred vessel dmax (maximum drop diameter) will be given

    bydmax

    T

    We0.6T (29)

    and

    WeT =cN

    2T3

    (30)

    where WeT is thestirred-tankWeber number and istheinterfacial

    tension.116 To account for the effect of volume fraction on dmax, a

    linear concentrationcorrection functionwiththe followinggeneral

    form was used:

    dmax

    T= c1(1+ c2)We

    0.6T (31)

    where c1 and c2 are constants and is the volume fraction of

    the dispersed phase.61 The constant c2 is considered equal to 3

    when it accounts for turbulence damping at low dispersed-phase

    concentrations, or higher than 3 for coalescing systems.117

    The maximum drop diameter is also a key parameter since it is

    generally considered proportional to the Sauter mean diameter

    d32, although this has been questioned.118 The Sauter mean

    diameter, commonly used in processes depending on interfacial

    area, is defined as the ratio of the third to the second moment of

    the drop size distribution:

    d32 =

    ki=1

    nid3i

    ki=1

    nid2i

    (32)

    where kis the number of size classes, ni the number of drops and

    di the diameter of drops in size class i.117119

    These aforementioned equations predict an increase in drop

    size with increasing dispersed phase volume fraction, up to

    about 40%. At higher fractions (usually above 50%), a further

    increasein thedispersedphase concentrationresultsin decreasing

    drop size. This behavior is attributed to a change in the dropbreakage mechanism, from turbulent eddy at low concentrations,

    to boundary layer at high concentrations.117,120

    BuchsandZoels120 showedthatshakeflasksledtolowerlevelsof

    hydro-mechanical stress, as power consumption was much more

    evenly distributed than in stirred tanks. This is due to different

    mixing mechanism present in either reactor. Thus, in stirred tank

    reactors, the power drawn from the region close to the stirrer is

    quite high comparedwith the average P/V,sinceasmallimpelleris

    used to promote mixingin a large vessel. a relatively small impeller

    agitates a relatively bulky tank, leading to higher power drawn

    in the region adjacent to the stirrer compared to the measured

    average or specific P/V. Onthe other hand, whenshakenflasks are

    used, power is more homogeneously distributed throughout thevessel, since energy is introduced through a relatively large wall

    area.121 These effects must be taken also into account if cultures

    are used where shear stress influences physiological behavior,

    such as filamentous organisms.

    Cull etal.116 exploited the use of two-phase systems for

    the whole-cell bioconversion of 1,3-dicyanobenzene (1,3-DCB)

    to 3-cyanobenzamide.with resting cells of Rhodococcus R312.

    Volumetric power consumption (or N3Di2 constant) and constant

    tip speed (or NDi constant) were the two criteria tested for scale-

    up, from a 3 L to a 75 L reactor. The system was composed

    of a phosphate buffer phase, where the cells were dispersed,

    and toluene, which contained 1,3-DCB and, throughout the

    bioconversion, the 3-cyanobenzamide formed.

    J Chem Technol Biotechnol2010; 85: 11841198 c 2010 Society of Chemical Industry www.interscience.wiley.com/jctb

  • 8/2/2019 2010_J Chem Tech Biotech_Bioprocess Scale-up

    11/15

    www.soci.org MPC Marques, JMS Cabral, P Fernandes

    Sauter mean drop diameters and drop size distributions were

    verysimilarforscale-uponthebasisofP/V, in geometrically similar

    reactors. In all cases the drop size distributions obtained were log-

    normal. The results presented in that work were obtained in

    a specific phase system comprising 20% v/v toluene dispersed

    in an aqueous buffer containing up to 10 gcell wet weight L1

    of Rhodococcus R312 cells. Despite being system specific, this

    work contributed to establish specific power consumption as a

    suitable tool for scale-up of two-phase bioconversion systems.

    Moreover, it is demonstrated that a scale increase of 25-fold is

    possible maintaining system specificity. The use of smaller reactors

    leads to cost reduction with reagents and overall process power

    consumption, among others. Moreover, problems that might be

    encountered at the large scale can be rapidly and efficiently

    identified in smaller scale.

    Another example of scale-up of a two-phase system based on

    the volumetric power consumption is the production 6-pentyl-

    -pyrone (6PP) performed by cells of Trichoderma harzianum.

    Rocha-Valadez etal.121 reported a 20-fold scale-up, from 500 mL

    shake flasks to a 10 L stirred tank reactor. 6PP production

    followed a sigmoid-shaped relationship with P/V, regardless of

    the production scale or impeller diameter. Synthesis of 6PP wastriggered earlier in the stirred reactor, suggesting a physiological

    response mechanism of T. harzianum towards hydrodynamic

    stress. Overall, at low P/V values (from 0.08 to 0.4 kW m3) a

    gradual increase in P/V improved 6PP production; however, for

    P/V > 0.6 kW m3, 6PP concentration was significantly reduced

    due to higher hydrodynamic stress.

    P/Vshiftinfluenced microbialgrowth in the stirred reactor since

    the specific growth rate decreased from 0.052 h1 at 0.08 kWm3

    to 0.033 h1 at 1.6 kWm3. On the other hand the specific growth

    rate was practically unaffected by P/V in the shake flask system.

    Higher shear rates involve drastic physiological changes including

    cellular differentiation and conidiospore production. This work

    showed thatphysiological changes observedduring process scale-

    up (i.e. microbial growth, production rates and sporulation) weretriggeredas resultof differences in the shear conditions prevailing

    in the two systems employed.

    P/V provides a suitable scale-up criterion from miniaturized

    devices when two-phase systems are involved. Further studies

    would be welcome addressing particular cases where MWP are

    used as the starting platforms.

    OTHER SCALE-UP PARAMETERSExamplesoftheapplicationofotherscale-upparametersarescarce

    due to the fact that both kLa and P/V incorporate information

    contained in the Reynolds number and mixing time, among

    others. Nonetheless, examples of such criteria are given.

    Constant impeller tip speed

    The impeller tip speed, vtip, is expressed as

    vtip = NT (33)

    Tip speed is used as a rule for scale-up when the relationship

    between shear and morphology is far from well understood, as

    happens in mycelial cultures.122 A rough rule of thumb suggests

    that cell damage can occur at tip speeds above 3.2 m s1, but

    the exact value is influenced by many factors such as broth

    rheology. Calculated tip speeds are usually greater than 3 m s1

    for production scale reactors.122 Although useful for estimating

    the potential for hyphae breakage and thus alteration of broth

    morphology when branched yeast, filamentous bacterial and

    fungal fermentations are involved, tip speed is less useful for

    single cell bacterial or yeast fermentations. If scale-up is carried

    out using constant tip speed (with geometric similarity), then the

    value ofP/V is often lowered, which can adversely affect aeration

    efficiency. It is possible to overcome this drawback by using more

    impellers in the larger vessel in such a way that both tip speed and

    P/Vare kept constant. Tip speed influences impeller shear, which

    is proportional to the product of impeller tip speed and impeller

    diameter, NDi2, for turbulent flow conditions.122

    Hiruta etal.123 demonstrated that maintaining impeller tip

    speed of 270 m min1 allowed scaling-up the production of -

    linolenic acid by Mortierella ramanniana mutant MM 15-1 from a

    30 L to a 1 m3 reactor.

    More recently, Dubey etal.124 scaled-up the demethylation

    of colchicine and their derivatives using Bacillus megaterium

    ACBT03 cells, from a 5 L to a 70 L reactor. Under optimum culture

    conditions the key monitoring factors to scale-up the process of

    demethylation were aeration rate of 2.5 vvm and impeller tip

    velocity of 4710 cm min1

    .

    Similar Reynolds number

    Reynolds number is expressed as:

    Re =TN2

    (34)

    The use of constant Reynolds number is hardly ever used

    for fermentation scale-up, since the effect of aeration on the

    process is notincorporated,125 andthe Reynoldsnumber generally

    increases for successful scale-up designs. Other dimensionless

    groups have also been examined for scale-up with limited

    success, often resulting in technically unrealistic equipment andoperatingparameters. As it is difficultto maintain all dimensionless

    parameters constant upon scale-up, those most important to the

    process must be identified accurately.122

    Constant mixing time

    The mixing time tm denotes the time required for the reactor

    composition to achieve a specified levelof homogeneity following

    additionof a tracerpulse at a singlepointin thevessel. Mixingtime

    containsinformation about flowand mixing within the reactorand

    can be useful forbiosynthesis processes scale-up. Themixing time

    tm is defined as1

    tm =V

    NfNT3

    (35)

    where Nf is the pumping number. The mixing time is typically

    measured in stirred vessels. Nonetheless, there is an increased

    interest in determining this parameter in shaken reactors. Gerson

    etal.126 used a mixing probe to determine fluid mixing in a 1 L

    shaken flask with filling volume of 540 mL at different shaking

    frequencies. The results demonstrated that it is possible to

    make such measurements, and that the mixing intensity rises

    monotonically with shaking frequency in both stirred reactors and

    shake flasks. Also, the range of mixing intensities measured by the

    device is similar in both systems over the range studied.

    Recently, Nealon etal.127 developed a high-speed video

    technique for the accurate quantification of jet macro-mixing

    times in static microwell plates, which also enables visualisation

    www.interscience.wiley.com/jctb c 2010 Society of Chemical Industry J Chem Technol Biotechnol2010; 85: 11841198

  • 8/2/2019 2010_J Chem Tech Biotech_Bioprocess Scale-up

    12/15

    Bioprocess scale-up: parameters to be used www.soci.org

    of jet formation and liquid flow patterns within the wells. Three

    microwell geometries were investigated: a single well from a

    standard 96-round well plate and dimension modified 96-well

    plate. A general correlation was found for the time to reach 95%

    homogeneity (s):

    t95 =2.60D1.5h0.5

    u0dn(36)

    where dn is the nozzle diameter, D is the well or vessel diameter, his the liquid height and u0 is the nozzle velocity.

    CONCLUSIONSDespite the relevance of the scale-up issue in biotechnology,there

    is no straightforward and uniform strategy to tackle this matter.

    A suitable scale-up criterion is elaborated in accordance with the

    individual product, process and the facility.

    An overall scale-up strategy consists of (i) a comprehensive

    and detailed process characterization, and (ii) an appropriate

    process control and process design. Owing to the complex

    nature of bioprocesses (biotransformation, bioconversion and

    fermentation) successful scale-up in most cases will not be basedon a straightforward strategy, but rather is the outcome of an

    independent optimization on each process scale. Nonetheless,

    bioconversions and fermentations are tentatively scaled-up on

    the basis of kLa or volumetric power consumption from lab- to

    pilot-scale. With the increasing ability to incorporate in MWPs and

    in miniature and microreactors devices that allow the monitoring

    of key process variables, such as dissolved oxygen tension,

    pH, optical density or protein production, the application of

    such small scale reactors is gaining widespread use in process

    optimization, rather than just in the early stages of screening.

    This pattern, that places small scale reactors in later stages of

    process development, puts more focus on the need for increasing

    the accuracy of scaling criteria. This has clearly been taken into

    consideration, as shown by the number of papers dedicatedto this matter that have been published recently. Accordingly

    the number of case studies reporting the successful scaling

    up from MWPs and similar devices to bench scale reactor

    or above is increasing. The insight required to obtain more

    knowledge on this matter has received significant contributions

    from the developments in image analysis, data processing and

    development of predictive models, namely through the use

    of CFD. In spite of the many existing studies, no examples of

    direct scale translation to industrial scale exist. There is a strong

    likelihood that scaling from shaken vessels to plant scale will

    not become a reality, given distinct hydrodynamic environments,

    preventing fullreproducibility, hencecompromisingthe outcome.

    The translation from miniaturized to production environment is

    nevertheless possible by scaling-out/numbering-up rather than

    scaling-up. In this approach, large numbers of microreactors,

    operating in continuous mode, are assembled so that production

    up to tons/year basis can be achieved.

    ACKNOWLEDGEMENTSMPC Marques and P Fernandes thank Fundacao para a Ciencia

    e Tecnologia (Portugal) for financial support in the form of a

    PhD grant SFRH/BD/24433/2005 and programme Ciencia 2007,

    respectively. This work was partially funded by research project

    POCI/SAU-MMO/59370/2004 from Fundacao para a Ciencia e a

    Tecnologia (Portugal).

    REFERENCES1 Schmidt FR, Optimization and scale up of industrial fermentation

    processes.Appl Microbiol Biotechnol68:425 435 (2005).2 Najafpour GD, Biochemical Engineering and Biotechnology. Elsevier

    Science, Amsterdam, The Netherlands (2007).3 Fernandes P and Cabral JMS, Microlitre/millilitre shaken bioreactors

    in fermentative and biotransformation processes a review.Biocatal Biotransform 24:237 252 (2006).

    4 Betts JI and Baganz F, Miniature bioreactors: current practices and

    future opportunities. Microbial Cell Factories 5:21 (2006).5 Marques MPC, Magalhaes S, Cabral JMS and Fernandes P,Characterizationof 24-well microtiterplate reactors fora complexmulti-step bioconversion: from sitosterol to androstenedione.

    J Biotechnol141:556 561 (2009).6 Micheletti M, Barrett T, Doig SD, Baganz F, Levy MS, Woodley JM,

    etal, Fluid mixing in shaken bioreactors: implications for scale-uppredictions from microlitre-scale microbial and mammalian cellcultures. Chem EngSci61:29392949 (2006).

    7 Gill NK, Appleton M, Baganz F and Lye GJ, Quantification ofpower consumption and oxygen transfer characteristics of astirred miniature bioreactor for predictive fermentation scale-up.Biotechnol Bioeng 100:11441155 (2008).

    8 Zhang H, Lamping SR, Pickering SCR, Lye GJ and Shamlou PA,Engineering characterisation of a single well from 24-well and96-well microtiter plates. Biochem Eng J40:138 149 (2008).

    9 Micheletti M and Lye GJ, Microscale bioprocess optimization. Curr

    Opin Biotechnol17:611 618 (2006).10 Weuster-Botz D, Puskeiler R, Kusterer A, Kaufmann K, John GT and

    Arnold M,Methodsandmilliliterscaledevicesforhigh-throughputbioprocess design. Bioprocess Biosyst Eng 28:109 119 (2005).

    11 Marques MPC, Cabral JMS and Fernandes P, High throughputin biotechnology: from shake-flasks to fully instrumentedmicrofermentors. Recent Patents Biotechnol3:124 140 (2009).

    12 Burke F, Scale up and scale down of fermentation processes. inPracticalFermentation Technology, ed.by McNeil B andHarvey LM.John Wiley & Sons, pp. 231 269 (2008).

    13 Garcia-Ochoa F and Gomez E, Bioreactor scale-up and oxygentransfer rate in microbial processes: an overview. Biotechnol Adv27:153 176 (2009).

    14 Buchs J,Introductiontoadvantagesandproblemsofshakencultures.Biochem EngJ7:198 (2001).

    15 Liu YS, Wu JY and Ho KP, Characterization of oxygen transferconditions and their effects on Phaffia rhodozyma growth andcarotenoid production in shake-flask cultures. Biochem Eng J27:331 335 (2006).

    16 Doran P, Bioprocess Engineering Principles. Academic Press, London(1995).

    17 Calik P, Yilgor P, Ayhan P and Demir AS, Oxygen transfer effectson recombinant benzaldehyde lyase production. Chem Eng Sci59:50755083 (2004).

    18 Gomez E, Santos VE, Alcon A and Garcia-Ochoa F, Oxygen transportrate on Rhodococcus erythropolis cultures: effect on growth andBDS capability. Chem Eng Sci61:45954604 (2006).

    19 Ju L-K andSundarajan A, Modelanalysisof biologicaloxygentransferenhancement in surface-aerated bioreactors. Biotechnol Bioeng40:13431352 (1992).

    20 Islam RS, Tisi D, Levy MS and Lye GJ, Scale-up of Escherichia coligrowth and recombinant protein expression conditions frommicrowell to laboratory and pilot scale based on matched kLa.

    Biotechnol Bioeng 99:128 1139 (2008).21 Kostov Y, Harms P, Randers-Eichhorn L and Rao G, Low-costmicrobioreactor for high-throughput bioprocessing. BiotechnolBioeng 72:346 352 (2001).

    22 Zanzotto A, Szita N, BoccazziP, Lessard P, Sinskey AJ andJensen KF, Membrane-aerated microbioreactor for high-throughput bioprocessing. Biotechnol Bioeng 87:243 254 (2004).

    23 Betts J, Doig S and Baganz F, Characterization and application ofa miniature 10 mL stirred-tank bioreactor, showing scale-downequivalence with a conventional 7 L reactor. Biotechnol Prog22:681 688 (2006).

    24 Kensy F,Engelbrecht C and Buchs J, Scale-up from microtiterplatetolaboratory fermenter: evaluation by onlinemonitoringtechniquesof growthand proteinexpressionin Escherichiacoliand Hansenula

    polymorpha fermentations. Microb Cell Factories 8:68 (2009).25 Doig SD, Ortiz-Ochoa K, Ward JM and Baganz F, Characterization

    of oxygen transfer in miniature and lab-scale bubble column

    J Chem Technol Biotechnol2010; 85: 11841198 c 2010 Society of Chemical Industry www.interscience.wiley.com/jctb

  • 8/2/2019 2010_J Chem Tech Biotech_Bioprocess Scale-up

    13/15

    www.soci.org MPC Marques, JMS Cabral, P Fernandes

    bioreactors and comparison of microbial growth performancebased on constant k(L)a. Biotechnol Prog 21:11751182 (2005).

    26 Vant Riet K, Review of measuring methods and nonviscousgasliquid mass transfer in stirred vessels. Ind Eng Chem ProcessDesign Dev18:357 364 (1979).

    27 Duetz WA, Microtiter plates as mini-bioreactors: miniaturization offermentation methods. Trends Microbiol15:469 475 (2007).

    28 Anderlei T and Buchs J, Device for sterile online measurement of theoxygen transfer rate in shaking flasks. Biochem Eng J 7:157162

    (2001).29 Anderlei T, Zang W, Papaspyrou M and Buchs J, Online respiration

    activity measurement (OTR, CTR, RQ) in shake flasks. Biochem EngJ17:187 194 (2001).

    30 GuezJS, Muller CH, Danze PM, Buchs J and Jacques P, Respirationactivity monitoring system (RAMOS), an efficient tool to studythe influence of the oxygen transfer rate on the synthesisof lipopeptide by Bacillus subtilis ATCC6633. J Biotechnol134:121 126 (2008).

    31 Wittmann C, Kim HM, John G and Heinzle E, Characterization andapplication of an optical sensor for quantification of dissolved O2in shake-flasks. Biotechnol Lett25:377 380 (2003).

    32 Zhang Z, Perozziello G, Boccazzi P, Sinskey AJ, Geschke O andJensen KF, Microbioreactors for Bioprocess Development. JALA12:143 151. (2007).

    33 Kensy F, Zimmermann HF, KnabbenI, Anderlei T, Trauthwein H,

    Dingerdissen U, etal, Oxygen transfer phenomena in 48-wellmicrotiter plates: determination by optical monitoring of sulfiteoxidation and verification by real-time measurement duringmicrobial growth. Biotechnol Bioeng 89:698 708. (2005).

    34 PuskeilerR, Kaufmann K and Weuster-Botz D, Development,parallelization, and automation of a gas-inducing milliliter-scale bioreactor for high-throughput bioprocess design (HTBD).Biotechnol Bioeng 89:512 523 (2005).

    35 John GT, Klimant I, Wittmann C and Heinzle E, Integrated opticalsensing of dissolved oxygen in microtiter plates: A novel tool formicrobial cultivation. Biotechnol Bioeng 81:829 836 (2003).

    36 Kensy F, Zang E, Faulhammer C, Tan R-K and Buchs J, Validationof a high-throughput fermentation system based on onlinemonitoring of biomass and fluorescence in continuously shakenmicrotiter plates. Microb Cell Fact8:31 (2009).

    37 Schneider K, Schutz V, John GT and Heinzle E, Optical device forparallel onlinemeasurement of dissolved oxygen andpH in shakeflask cultures. Bioprocess Biosyst Eng. DOI 10.1007/s00449-009-0367-0 (E-pub ahead of print). Latest info required.