You are on page 1of 8

Power Consumption in Shaking Flasks

on Rotary Shaking Machines:


II. Nondimensional Description of
Specific Power Consumption and
Flow Regimes in Unbaffled Flasks at
Elevated Liquid Viscosity

Jochen Büchs,1 Ulrike Maier,1 Claudia Milbradt,1 Bernd Zoels2


1
Chair of Biochemical Engineering, Sammelbau Biologie, Aachen
University of Technology, D-52074 Aachen, Germany, telephone:
+49-241-805546; fax: +49-241-8888-265; e-mail:
buechs@biovt.rwth-aachen.de
2
ZET/ZT-A15, BASF AG, Ludwigshafen, Germany
Received 6 September 1999; accepted 14 February 2000

Abstract: This article is the second part of a series pre- larger companies working in the field of biotechnology, up
senting and modeling the hydrodynamics and specific to several hundred-thousand experiments are carried out us-
power consumption in shaking flasks on rotary (orbital)
ing shaking flasks annually. These bioreactors are well
shaking machines. In part I, a new method was intro-
duced that enables the accurate determination of the suited to more or less routine tasks like medium evaluation
specific power consumption in shaking flasks. The and strain selection, involving the need for many replicate
method was first applied to investigate unbaffled flasks experiments. Nowadays, the design of agitated tank biore-
with a nominal volume of ⱕ1 L at low viscosity. In part II, actors with a standard geometry no longer presents a big
the results for the specific power consumption of un-
baffled shaking flasks at elevated viscosities are investi-
problem. In contrast, shaking flasks, in which the majority
gated after varying shaking frequency, flask size, filling of biotechnological development takes place, are only in-
volume, and shaking diameter. The theory introduced in sufficiently described.
part I is extended to liquids of elevated viscosities using One of the important parameters in fermentation of aero-
nondimensional equations. With these results, the spe- bic microorganisms in shaking flasks and agitated tank bio-
cific power consumption in unbaffled shaking flasks can
now be fully described. For the first time, the phenom-
reactors is the specific power consumption (per unit vol-
enon of the liquid being “out of phase” is observed and ume). In part I of this series (Büchs et al., 2000) a new
described. This occurs at certain operating conditions method was introduced that enables the accurate determi-
and is characterized by an increasing amount of liquid nation of the specific power consumption in a shaking flask
not following the movement of the shaking table, thus on rotary (orbital) shaking machines. The aim of part II is to
reducing the specific power consumption. This, of
course, has much relevance for practical work with mi-
present algebraic equations based on simple, physically
crobial cultures. The phenomenon of being “out-of- sound, nondimensional models with which the specific
phase” is described in the form of a newly defined non- power consumption and reasonable operating conditions for
dimensional phase number (Ph) in analogy to a partially shaking flasks may be calculated for different flask sizes at
filled, rotating horizontal drum. The Ph can be used to elevated viscosities for newtonian fluids (up to ␩ ⳱ 200
determine reasonable operating conditions for shaking
mPa ⭈ s). An investigation of the influence of baffles on
flask experiments when using viscous media, avoiding
unfavorable “out-of-phase” operation. © 2000 John Wiley “out-of-phase” operating conditions will be presented in a
& Sons, Inc. Biotechnol Bioeng 68: 594–601, 2000. subsequent article of this series.
Keywords: power consumption; shaking flasks, model-
ing; viscous media; operating conditions; filamentous
microorganisms MODEL DEVELOPMENT
It has been shown that the specific power consumption in
shaking flasks at low viscosity may generally be described
INTRODUCTION
by the modified power number (Ne⬘), which is a function of
Due to their simplicity shaking flasks are a major tool for the flask Reynolds number (Re), according to (Büchs et al.,
parallel studies in microbiological process development. In 2000):
P
Ne⬘ = = C ⭈ f 共Re兲 (1)
Correspondence to: J. Büchs ␳ ⭈ n ⭈ d 4 ⭈ VL1 Ⲑ 3
3

© 2000 John Wiley & Sons, Inc.


␳ ⭈ n ⭈ d2 (1982) observed a bistable region. In a recent and more
Re = (2) detailed examination of horizontal drums of finite length it

was shown that the side faces influence the flow signifi-
This form of equation can generally be used for operating cantly. This results in highly complicated hydrodynamical
conditions at which the bulk of the liquid within the flask interactions and many different flow regimes (Thoroddsen
circulates “in-phase” with the shaking table. and Mahadevan, 1997). It can be expected that the situation
In contrast, at certain operating conditions (with the vis- in shaking flasks is just as complicated. Nevertheless, the
cosity being the main influencing factor), the bulk of the simplifying approaches of Deiber and Cerro (1976) and Se-
liquid remains at the base of the flask and shows only very mena and Khmelëv (1982) are quite helpful and may serve
little relative movement. These operating conditions are as a sufficient theoretical background for our ultimate goal
termed as “out-of-phase” throughout this study. in the present work.
To systematically determine these unsuitable out-of- To transfer these conditions to shaking flasks and to un-
phase operating conditions an analogy to a partially filled, derstand the accelerations present, it is helpful to divide the
rotating horizontal drum is used. The ratio of forces acting overall liquid motion into two partial movements: the rotary
in the plane perpendicular to the shaking flask axis is similar movement of the liquid relative to the flask, and the trans-
to that present in a horizontal drum. Yet, the force present in latory movement of the complete shaking system (including
the shaking flask in the radial direction is the centrifugal the individual flasks) relative to the ground. The liquid ac-
force induced by the translatory motion of the shaking table. celeration relative to the flask for “in-phase” conditions is
In the case of the shaking flask the gravitational force acts simplified to being equal to (2 ⭈ ␲ ⭈ n)2 ⭈ d/2 (assuming d ⳱
in the axial direction. The flow regimes present in a hori- constant). The radial acceleration of the translatory moving
zontal drum have been investigated in detail by Deiber and system relative to the ground is equal to (2 ⭈ ␲ ⭈ n)2 ⭈ do/2.
Cerro (1976) as well as Semena and Khmelëv (1982). These Hence, the radial Froude (Frr) number (radial/radial forces)
investigators assumed that the drum is infinitely long, so in shaking flasks is equal to d/do. As this expression consists
that the side faces have no influence on the flow behavior. of geometric dimensions only, it is defined as the geometry
They showed that the flow regime is described by a radial ratio (GR):
Froude (Frr) number (radial/radial forces) and a liquid film
Reynolds (Ref) number: d
GR = (7)
do
共2 ⭈ ␲ ⭈ n兲 ⭈ d2
Frr = (3) Our modeling approach based on the GR is reasonable only
2⭈g
for elevated shaking frequencies, at which nearly all liquid
␳ ⭈ 共2 ⭈ ␲ ⭈ n兲 ⭈ f 2 is rotating on the flask walls. These are the conditions,
Ref = (4) however, that are most commonly used in practical biotech-

nological development work. At low shaking frequencies,
It should be noted that the liquid film Reynolds number the gravitational acceleration, which acts in the axial direc-
(Ref) is not identical to the flask Reynolds number (Re) tion of the shaking flask, is predominant over the centrifugal
introduced in Eq. (2). As illustrated in Figure 1, the symbol acceleration and the bulk of the fluid moves on the base of
f represents the liquid film thickness for the hypothetical
case of a uniform distribution of the liquid inside of a hori-
zontal drum with the length h, according to:

VL = ⭈ 共d 2 − d 2i 兲 ⭈ h (5)
4
Hence, the liquid film thickness ( f ) is given by:
共d − di兲
f= (6)
2
Two distinct flow regimes exist as shown in Figure 1 (sepa-
rated by boundary lines I and II or III). For high radial
Froude (Frr) or low liquid film Reynolds (Ref) numbers
(high viscosity, low filling volume) the fluid is almost com-
pletely distributed on the whole drum wall (regime A). For
low radial Froude or high liquid film Reynolds numbers
(low viscosity, large filling volume) only a very thin liquid Figure 1. Main flow conditions inside a horizontal rotating drum
(adapted from Semena and Khmelëv, 1982). (A) Complete distribution of
film is distributed on the drum wall, with the majority of the the liquid on the drum wall. (B) Region of entrained thin liquid layer.
liquid remaining in a groove at the base of the drum (regime Boundary line I: Frr ⳱ 2.2 ⭈ Re0.77
f ; boundary line II: Frr ⳱ 2.2 ⭈ Re0.13
f ;
B). Between boundary lines II and III Semena and Khmelëv boundary line III: Frr ⳱ 2.2 ⭈ Ref0.44.

BÜCHS ET AL.: POWER CONSUMPTION IN SHAKING FLASKS ON ROTARY SHAKING MACHINES. II 595
the shaking flask. This resembles the case of the horizontal mPa ⭈ s were prepared and used throughout all experiments.
drum of finite length, treated earlier, where the side walls To ensure near-constant viscosity, the solutions contained
influence the hydrodynamics. This is not covered by our 0.5 M sodium chloride and were stored for at least 7 days
model. before use. Before each experiment the viscosity was
To distinguish between operating conditions (shaking fre- checked again and readjusted if necessary. Between each
quencies) that are or are not covered by our model, an axial experiment the flasks were thoroughly washed out with
Froude (Fra) number (radial/axial forces) for shaking flasks deionized water and dried at 60°C. The measurements of
is defined as follows: viscosity and power consumption were taken at a tempera-
ture of 37°C.
共2 ⭈ ␲ ⭈ n兲2 ⭈ do
Fra = (8)
2⭈g
RESULTS AND DISCUSSION
To determine the liquid film Reynolds number (Ref) for the
shaking flask in an analogous form to the horizontal drum, “In-Phase” Operating Conditions
the liquid film thickness (f) for the uniform distribution of
To present the results of the 2143 measuring points for
the liquid on the flask wall must be defined. As with the
shaking flasks in a closed form, the modified Newton num-
drum geometry, the distribution of liquid is approximated
ber (Ne⬘) is plotted over the flask Reynolds number (Re) in
by a hollow liquid cylinder. We therefore use the maximum
Figure 2.
inside flask diameter, d, to describe the outer diameter of the
The “in-phase” operating conditions (1115 individual
hollow liquid cylinder; the inner diameter is expressed by di
points) for flask sizes of 100 to 2000 mL are represented by
(analogous to Fig. 1). Eq. (6) is also applied to calculate the
the measuring points concentrated in the upper limit of the
film thickness (f).
diagram. This limit shows the same course as the power
To determine di, the simplifying relation h ≅ VL1/3 is used
number (Ne) for unbaffled agitated tanks in the transition
as a characteristic length scale for the height of the friction
range between laminar and turbulent flow. When working
area between the rotating liquid and the flask wall (height of
with liquids of waterlike viscosity in shaking flasks, the
the assumed hollow liquid cylinder) as in the derivation of
Reynolds numbers indicate (Re > 104) that the flow regime
the modified power number, Ne⬘ (Büchs et al., 2000). By
is only just turbulent (Büchs et al., 2000). Fluids of elevated
rearranging Eq. (5) to di and inserting this expression into
viscosity lie in the transition range to laminar flow (see
Eq. (6), the liquid film thickness (f) is described by:
Fig. 2).
d
f= ⭈
2
冉 冑 冉 冊冊
1− 1−
4 VL1 Ⲑ 3
␲ d
2

(9)
These “in-phase” points are fitted using a least-square-
error method resulting in the following equation:
Ne⬘ ⳱ 70 ⭈ Re−1 + 25 ⭈ Re−0.6 + 1.5 ⭈ Re−0.2 (11)
The liquid film Reynolds number (Ref) for a shaking flask is
−1
now obtained by replacing f in Eq. (4) by Eq. (9): This correlation consists of a laminar (Re ), a transition

冉 冑 冉 冊冊
(Re−0.6), and a turbulent term (Re−0.2) analogous to well-
2 2
␳ ⭈ 共2 ⭈ ␲ ⭈ n兲 d 2 4 VL1 Ⲑ 3 known hydrodynamic problems (Schlichting, 1979). Using
Ref = ⭈ 1− 1− the fitted curve Eq. (11) and the definition of the modified
␩ 4 ␲ d
Newton number (Ne⬘) according to Eq. (1), the specific

冉 冑 冉 冊冊 2 2 power consumption of favorable operating conditions (in-


␲ 4 VL1 Ⲑ 3
= Re ⭈ 1− 1− (10) phase system) can now be determined in advance.
2 ␲ d
“Out-of-Phase” Operating Conditions
MATERIALS AND METHODS
A large proportion of the measuring points (1028 points) in
Measurements (a total of 2143 operating points) were taken Figure 2 lies in a region well below the fitted curve. These
using the measuring method described in part I of this series points represent unfavorable operating conditions (“out-of-
(Büchs et al., 2000). The flask size (100 to 2000 mL, DIN phase” system).
12380), the shaking diameter of the shaker drive (do ⳱ 2.5 To demonstrate “out-of-phase” operating conditions, an
and 5 cm), the viscosity (␩ ⳱ 0.8 to 200 mPa ⭈ s), the example is given in Figure 3. The modified Newton number
shaking frequency (n ⳱ 80 to 380 min−1 L/min), and the (Ne⬘) is depicted over the flask Reynolds number (Re) for a
filling volume (4% to 20% of the nominal flask volume) 250-mL flask, a filling volume of 25 mL and a shaking
were varied. The viscosity was varied using aqueous solu- diameter of 2.5 cm for varying viscosities. Up to a viscosity
tions of polyvinylpyrrolidone (Luviskol, BASF AG, Lud- of 4 mPa ⭈ s, the measuring points are well represented by
wigshafen, Germany) in concentrations of 3 to 300 g/L. the curve calculated by Eq. (11). Starting from a viscosity of
These solutions show nearly newtonian flow behavior. 17 mPa ⭈ s the difference between the measured and the
Stock solutions of approximately 4, 17, 33, 60, 100, and 200 calculated “in-phase” [Eq. (11)] modified Newton number

596 BIOTECHNOLOGY AND BIOENGINEERING, VOL. 68, NO. 6, JUNE 20, 2000
Figure 2. Closed presentation of all measuring points (2143) for unbaffled shaking flasks in form of the modified Newton number (Ne⬘) dependent on
the flask Reynolds number (Re) with variation of flask size, viscosity, shaking diameter, and shaking frequency. The lower boundary curve [Eq. (12)] is
used for the model development (see text). The final differentiation of “in-phase” (large symbols) and “out-of-phase” (small cross symbols) operating
conditions is determined by Eqs. (8), (10), and (14).

(Ne⬘) increases continually, indicating the onset of “out-of-


phase” operating conditions.
The photographs shown in Figure 4 show the liquid dis-
tribution in the flasks for three representative operating con-
ditions. At a viscosity of 1 mPa ⭈ s, the bulk of the liquid is
at all times driven in the direction of the centrifugal accel-
eration inside the flask (Fig. 4a). At increasing viscosities
the bulk of the liquid shows an increasing delay (of up to
90°) relative to the direction of the centrifugal acceleration
(Fig. 4b). At a viscosity of 35 mPa ⭈ s (Fig. 4c), the liquid is
completely “out-of-phase”: the majority of the liquid re-
mains at the base of the flask and the maximum liquid
height is significantly reduced. With this only a minor frac-
tion of the liquid is actually rotating on the flask wall and
the total momentum transfer (i.e., power consumption) is
thereby greatly reduced. A comparable phenomenon is well
known when operating magnetic stirrers. With these devices Figure 3. Modified Newton number (Ne⬘) over the flask Reynolds num-
the magnetic forces normally assure that the magnetic bar ber (Re) for a 250-mL flask with a shaking diameter of 2.5 cm and a filling
rotates “in-phase” with the magnetic drive. Yet, if the vis- volume of 25 mL for varying viscosities.

BÜCHS ET AL.: POWER CONSUMPTION IN SHAKING FLASKS ON ROTARY SHAKING MACHINES. II 597
cous forces of the liquid exceed the magnetic forces, the
stirrer bar will jump out of position and solely vibrate with-
out rotating. This results in near-zero power consumption.
As the complex flow conditions in the shaking flask are
somewhat simplified in our modeling approach, the in-
phase points do not lie perfectly on the fitted curve [Eq.
(11)] of Figure 2 but are scattered within a band. To account
for these scattered points, a lower boundary curve is intro-
duced in Figure 2, according to Eq. (12) (for derivation see
Appendix).

Figure 4. Liquid distribution in shaking flasks for a 250-mL flask with a


Ne⬘trans =
冉 70

Re ⭈ 10c Ⲑ
公1 + 1 Ⲑ m2
+
冉Re ⭈ 10 公 cⲐ
25
1+1 Ⲑ m2 冊 0.6


shaking diameter of 2.5 cm, a filling volume of 25 mL, and a shaking
frequency of 200 L/min for: (a) a viscosity of 1 mPa ⭈ s, Ph ⳱ 2.61; (b) a
+
1.5
⭈ 10−c Ⲑ
公m2+1 (12)
viscosity of 75 mPa ⭈ s, Ph ⳱ 0.83; and (c) a viscosity of 135 mPa ⭈ s, Ph
⳱ 0.59. First row: photographs taken in the direction of the centrifugal
acceleration. Second row: photographs taken perpendicular to the centrifu-
冉Re ⭈ 10 公
cⲐ 1+1 Ⲑ m2冊0.2

gal acceleration. The arrows on the left side of the photographs show the
direction of the centrifugal acceleration. Every measuring point that lies above the lower boundary
curve according to Eq. (12) is used in the following model
development (Fig. 5) as an “in-phase” operating point. It

Figure 5. Separation of “in-phase” (large symbols) from “out-of-phase” (small filled circle symbols) operating conditions in dependency of the geometry
ratio (GR) and the liquid film Reynolds number (Ref).

598 BIOTECHNOLOGY AND BIOENGINEERING, VOL. 68, NO. 6, JUNE 20, 2000
should be noted that the conditions depicted in Figure 2 as The following general tendencies can be deduced from Eqs.
being “in-phase” (with large symbols) do not correspond to (8), (10), and (14). The flow in a shaking flask tends to be
this definition, but are the results of Eq. (14) (see later). The “in-phase” at:
measuring points taken for the lowest viscosity liquid (wa-
1. High shaking frequency (n).
ter, Re > 104) are always “in-phase,” independent of the
2. Low viscosity (␩).
operating conditions.
3. Large filling volume (VL).
To predict “in-phase” conditions, the model of a partially
4. Large shaking diameters (do).
filled rotating horizontal drum (described earlier) has been
5. Small maximum inside shaking flask diameters (d).
applied to shaking flasks [Eq. (7) and Eq. (10)]. To elimi-
nate measuring points of low shaking frequencies, where the
majority of the liquid is in contact with the flask base and CONCLUSION
where our model is invalid, only measuring points with For the first time it has become possible to use algebraic,
axial Froude numbers (Fra) of >0.4 [Eq. (8)] are considered physically based, nondimensional equations [Eq. (11) and
further. Eq. (14)] to determine sensible operating conditions for un-
The “in-phase” and “out-of-phase” operating points re- baffled shaking flasks and to estimate the achievable power
sulting from Eqs. (8) and (12) are shown in Figure 5, using consumption. When working with shaking flasks of ⱕ1 L
the geometry ratio (GR), Eq. (7), and the film Reynolds nominal volume, “out-of-phase” operating conditions are
number (Ref), Eq. (10). The small filled-circle symbols only observed at elevated liquid viscosities. While in culture
show “out-of-phase” operating points. All other symbols experiments with bacteria and yeasts of single-cell morphol-
represent in-phase operating points, at which the liquid ogy this should not present problems, caution should be
moves synchronously with the rotating movement of the exerted with filamentous microorganisms. In this case, the
drive. Figure 5 shows that in the shaking flasks evaluated phenomenon of “out-of-phase” operating conditions may be
there is a clear boundary between the two hydrodynamic of great practical relevance. Working under such conditions
regimes, with only a small percentage of measuring points will significantly reduce the oxygen transfer and mixing
(8.9%) not in agreement. Using a trial-and-error method in intensity (detailed investigations are currently being con-
a least-squares sense, the boundary curve between “in- ducted in our laboratory). If a screening program with fila-
phase” and “out-of-phase” points is given by: mentous microorganisms is unintentionally performed using
“out-of-phase” operating conditions, the whole screening
d will suffer from inadequate substrate supply to the cells.
GR = = 0.79 + 2.38 ⭈ log10共Ref兲 (13)
d0 Improved strains will hardly be recognizable. In addition,
these operating conditions will make it very likely that
It is quite remarkable that the slopes of the boundary curve
screening would be led in a entirely wrong direction. Strains
in the lower left and upper right-hand corner of Figure 5
may attract attention by increased product titers that are in
(shaking flasks) are nearly equivalent to boundary lines I
reality not due to increased specific productivity but are the
and II of Figure 1 (rotating drum), respectively. The de-
result of an altered morphology. These changes in morphol-
duced hydrodynamic regimes in shaking flasks developed in
ogy of the strains (shorter hyphae with fewer branches or
analogy to the basic problem of the rotating horizontal drum
even single-cell growth) may lead to lower viscosities, re-
can be considered legitimate.
sulting in a switch to “in-phase” operating conditions and
To obtain an easy-to-use algebraic criterion to determine
therefore sufficient substrate supply of the cells, higher
the boundary up to which the liquid remains “in-phase”, we
metabolic activity and product titers. These seemingly “im-
introduce a nondimensional phase number (Ph). By multi-
proved” strains will probably not have increased specific
plying Eq. (13) by a factor of 1.26 and rearranging, the
productivities. The benefit of such screening procedures for
following equation results:
a production process in a stirred-tank fermentor, assuring
d0 sufficient substrate supply, is quite questionable. The same
Ph = ⭈ 共1 + 3 ⭈ log10共Ref兲兲 ⬎ 1.26 (14) applies to media development using “out-of-phase” operat-
d
ing conditions. Those media compositions that lead to
Eq. (14) states that operating points at which Ph > 1.26 are higher viscosities by polymers or by enhanced filamentous
“in-phase”. These “in-phase” operating conditions are de- growth will result in impaired substrate supply to the cells
fined by Fra > 0.4 [see Eq. (8)], the liquid film Reynolds and lower product titers. Possibly, the same media would
number (Ref) [Eq. (10)], and the phase number Ph > 1.26 turn out to be superior at “in-phase” operating conditions
[Eq. (14)]. These criteria are used to differentiate between with a sufficient supply of substrates. In at least one indus-
“in-phase” and “out-of-phase” operating conditions in Fig- trial case there is very strong evidence that the problems just
ure 2. The calculated differentiation agrees reasonably well described have been encountered (unpublished results). It
for practical application with the differentiation represented can be expected that quite similar problems exist in many
by the lower boundary curve [Eq. (12)]. It should also be other laboratories.
noted that the phase numbers for the operating conditions To determine sensible operating conditions for viscous
shown in Figure 4a–c are given in the corresponding legend. fermentation broths, which show non-newtonian behavior,

BÜCHS ET AL.: POWER CONSUMPTION IN SHAKING FLASKS ON ROTARY SHAKING MACHINES. II 599
information about the level of the average shear rate must be The form of Eq. (11) is logarithmized:

冋 册
available. The influence of the operating conditions on the
average shear rate in shaking flasks is currently being in- 70 25 1.5
y = log x
+ 0.6x
+ (A1)
vestigated in our laboratory. In addition, the appearance of 10 10 100.2x
“out-of-phase” operating conditions at low viscosities (wa-
ter) in baffled shaking flasks and in large unbaffled shaking This curve is differentiated to determine the local slope:
flasks (ⱖ2 L nominal volume) will be reported on in a later
publication. 70 25 1.5
x
+ 0.6 ⭈ 0.6x
+ 0.2 ⭈
dy 10 10 100.2x
The experiments presented were performed at the life science =− =m (A2)
dx 70 25 1.5
laboratory of the BASF AG, Ludwigshafen, Germany. The au- + +
thors thank B. Mayer and the staff members of the biological 10x 100.6x 100.2x
pilot plant (ZHV/TB) for their technical assistance. We also
thank Schott Glaswerke, Mainz, Germany, for kindly supplying To determine the x- and y-translation along the normal to
the shaking flasks for the photographs.
the local slope, the following geometrical relations (Fig.
A1) are applied in each point of the fitted curve [Eq. (11)]:
NOMENCLATURE
m = aⲐb and c2 = a2 + b2 (A3)
A friction area (m2)
C, C1, C2, C3, C4, C5 constants (−)
resulting in:
c transition factor, in this case 0.1 (−)
di inner diameter of the liquid film inside a cylin-
c
der (m)
b= (A4)
d maximum inside shaking flask diameter, drum
diameter (m)
公m2 + 1
do shaking diameter (orbital shaker) (m)
f liquid film thickness inside a horizontal drum or The coordinates x* and y* (see Fig. A1) of the lower bound-
a shaking flask regime assuming total distribu- ary curve are thus related by:
tion of the liquid, Eq. (6) (m)
f(Re) function depending on flow (−) x* = x − a and y* = y − b

冋 册
Fra axial Froude number, Eq. (8) (−)
Frr radial Froude number, Eq. (3) (−) 70 25 1.5
g gravitational acceleration (m/s2) y* = log + + −b
GR geometry ratio for shaking flasks, Eq. (7) (−) 10共x*+a兲
共10 兲
共x*+a兲 0.6
共10 共x*+a兲 0.2

h drum length, characteristic height of friction (A5)
area (m)
m gradient of the upper limiting curve at each Transforming Eq. (A5) to the linear state results in:


point (−)
n shaking frequency, number of revolutions (s−1) 70 25
Ne⬘trans = +
冉Re ⭈ 10 公 冊
Ne Newton number (−)
Ne⬘ modified Newton number, Eq. (1) (−)
Re ⭈ 10c Ⲑ
公 1+1 Ⲑ m2 cⲐ 1+1 Ⲑ m2
0.6


Ne⬘trans modified lower boundary Newton number, Eq.
(12) (−)
+
1.5
⭈ 10−c Ⲑ
公m2+1
冉Re ⭈ 10 公 冊
P power (W)
0.2
Ph phase number, Eq. (14) (−) cⲐ 1+1 Ⲑ m2
r drum radius (m)
Re flask Reynolds number, Eq. (2) (−) (A6)
Ref liquid film Reynolds number, Eq. (4) (horizon-
tal drum) and Eq. (10) (flask) (−)
VL filling volume (−)

Greek symbols

␳ density of liquid (kg/m3)


␩ dynamic viscosity of fluid (mPa ⭈ s)

APPENDIX

Derivation of Lower Boundary Curve


The lower boundary curve was designed in the logarithmic
scale by connecting points, which were moved by a specific
distance along the normal to the local slope of the fitted Figure A1. Geometric relations applied to determine the lower boundary
curve [Eq. (11)]. This method is now explained. curve.

600 BIOTECHNOLOGY AND BIOENGINEERING, VOL. 68, NO. 6, JUNE 20, 2000
The width of the band is defined by the distance between the measurement in unbaffled flasks at low liquid viscosity. Biotechnol
upper and lower boundary curve. To account for the major- Bioeng 68:589–593.
Deiber JA, Cerro RL. 1976. Viscous flow with a free surface inside a
ity of “in-phase” operating points, the distance along the horizontal rotating drum. I. Hydrodynamics. Ind Eng Chem Fund
normal to the local slope (m) was chosen to be c ⳱ 0.1. 15:102–110.
Schlichting H. 1979. Boundary layer theory, 7th ed. New York: McGraw
Hill.
Semena MG, Khmelëv YA. 1982. Hydrodynamic regimes of a liquid in a
References smooth-walled rotating heat pipe. 1. J Eng Phys 43:1235–1242.
Thoroddsen ST, Mahadevan L. 1997. Experimental study of coating flows
Büchs J, Maier U, Milbradt C, Zoels B. 2000. Power consumption in in a partially-filled horizontally rotating cylinder. Exper Fluids 23:
shaking flasks on rotary shaking machines: I. Power consumption 1–13.

BÜCHS ET AL.: POWER CONSUMPTION IN SHAKING FLASKS ON ROTARY SHAKING MACHINES. II 601

You might also like