Download as pdf or txt
Download as pdf or txt
You are on page 1of 18

Biotechnol. Prog.

2007, 23, 767−784 767

REVIEW
Foam and Its Mitigation in Fermentation Systems
Beth Junker*
Fermentation Development and Operations, Merck Research Laboratories, P.O. Box 2000, Rahway, New Jersey 07065

Key aspects of foaming and its mitigation in fermentation systems are presented. Foam properties
and behavior, conditions that affect foaming, and consequences of foaming are discussed, followed
by methods to detect and prevent foam, both without and with the use of antifoam, and their
implications. Antifoams were catalogued according to their class (e.g., polyalkylene glycols,
silicone emulsions, etc.) to facilitate recognition of antifoams possessing similar base composi-
tions. Relatively few published studies directly comparing antifoams experimentally are available,
but those reports found only partially identify clear benefits/disadvantages of any one antifoam
type. Consequently, desired characteristics, trends in antifoam application, and chemical types
of antifoams are evaluated on the basis of a thorough review of available literature reports
describing a specific antifoam’s usage. Finally, examples of specific foaming situations taken
from both the literature and from actual experience in an industrial fermentation pilot plant are
examined for their agreement with expected behavior.

Contents a little about foams but few bother to know very much” (Gaden
and Kevorkian, 1956). Today, a similar lack of collective and
Introduction 767
systematic knowledge exists. Consequently, the purpose of this
Foam and Impact of Antifoams 767 review article is to summarize past and current information
Properties and Behavior of Foam 767 concerning foam formation and its mitigation in fermentation
Conditions That Affect Foaming 768 applications.
Consequences of Foam 770
Detection of Foam 771 Foam and Impact of Antifoams
Prevention of Foam without Adding 771 Properties and Behavior of Foam. Foam is the dispersion
Antifoam of a gas in a continuous liquid phase, and thus foam dispersions
Impact of Antifoams 772 possess bulk densities closer to that of a gas rather than a liquid
Chemical Antifoams (Defoamers) 773 (Vardar-Sukan, 1992). Foam is distinctly different from tradi-
Desired Characteristics 773 tional gas holdup for which the gas-to-liquid volume ratio is
General Trends in Application 775 smaller and bubbles are more spherical (Prins and van’t Riet,
1987). Foam bubbles are located on top of rather than within
Types of Antifoams 776
the broth, as is the case for gas holdup (van’t Riet and Tramper,
Specific Foaming Situations 780 1991). A general definition of foam, applicable to bioreactors,
Literature Reports 780 determines foam to occur when gas holdup in a gas-liquid
Qualitative Pilot-Scale Foam Observations 780 dispersion is greater than 90% (Schubert et al., 1993). Other
Summary 781 authors have quantified the gas content of foam to be in the
range of 60-90% (van’t Riet and Tramper, 1991), with a gas-
to-liquid volume ratio >1and usually >3 (Prins and van’t Riet,
Introduction 1987).
Foaming is labeled a “general nuisance” in fermentation Foam generation is autocatalytic in some cases, exhibiting a
(Weng et al., 1997) because it is necessary to adequately aerate balance between the forces creating and destroying it (Vardar-
broth while keeping foam formation under control (Vardar- Sukan, 1992). Foam layers are dynamic in nature, remaining
Sukan, 1992). During industrial fermentations, foam control is the same size when the number of bubbles bursting equals the
perceived as an “empirical art” (Vardar-Sukan, 1992). Very few number of bubbles arriving in the foam layer (Lee et al., 1993).
submerged fermentation processes are performed without the Two main types of foam are common in fermentations: (1)
use of either natural or synthetic antifoam agents (Viesturs et Froths, sea, or “bubble bath” foams are short-lived, transitory,
al., 1982). Over 50 years ago, it was stated that “everyone knows and unstable, containing a wide range of bubble sizes. Bubbles
are polyhedral and form a honeycomb structure (Prins and van’t
* To whom correspondence should be addressed. Tel: (732) 594-7010. Riet, 1987). These foams contain significant amounts of
Fax: (732) 594-7698. Email: beth_junker@merck.com. entrained liquid and thus possess low gas-to-liquid volume
10.1021/bp070032r CCC: $37.00 © 2007 American Chemical Society and American Institute of Chemical Engineers
Published on Web 06/13/2007
768 Biotechnol. Prog., 2007, Vol. 23, No. 4

ratios. (2) Stable foams, resembling “beer head,” are uniform, present in complex media or when protein precipitated during
long-lasting, and rigid. Bubbles are spherical or slightly el- sterilization is utilized, and specifically caused by some property
lipsoidal with diameters <2-3 mm (Lee et al., 1993). These of the fermentation medium, or (2) later in the fermentation,
foams are low in liquid content and thus have high gas-to-liquid for example, owing to protein release from cell autolysis or some
volume ratios (Gaden and Kevorkian, 1956). other property of the culture itself (Hastings, 1954; Schügerl,
Foam is stable if gas bubbles remain separated by thin liquid 1985; Stanbury and Whitaker, 1984; König et al., 1979; Su,
walls and do not coalesce (Gaden and Kevorkian, 1956). 1995). This behavior parallels that found with activated sludge
Drainage, the runoff of liquid between bubbles in foam, is foam, namely, a white, frothy foam during start-up (first 3-4
dependent on the liquid viscosity and density (Gaden and days), followed by a sticky viscous foam later on when the
Kevorkian, 1956). Bioprocess foams tend to be non-coalescing culture becomes nutrient-limited (Jenkins et al., 1993). It also
when the presence of surface-active agents stabilizes foams as parallels the difference between frothy foam in wine fermenta-
they form (Vardar-Sukan, 1992); in other cases foams are tions owing to high rates of CO2 evolution and stable, strong
unstable or metastable (Berovič, 1992). foam in beer (Edwards et al., 1982). Foaming later in the
Foam stability is determined by the number of lamellae and fermentation process has been found easier to control using
the angles between them; foam is stable if three lamellae are antifoams (Hastings, 1954).
present at an angle of 120 degrees (Ghildyal et al., 1988). Foam There are five distinctive patterns of foam behavior during
stability also is favored if the surface tension of the gas-liquid fermentation (Stanbury and Whitaker, 1984 from Hall et al.,
system is less that of the pure solvent of solution (i.e., the liquid 1973): the foam (1) remains at a constant level throughout the
phase) (Ghildyal et al., 1988). The surface tension of pure water fermentation, (2) undergoes a steady fall during the initial
is approximately 72 dyn/cm (Hall, 1971), whereas the surface fermentation phase and then remains constant, (3) falls slightly
tensions of most fermentation media range between 60 and 65 during the initial phase and then rises, (4) demonstrates low
dyn/cm (Viesturs et al., 1982). The surface tension of water (or initial foaming and then rises, or (5) exhibits a combination of
media) is further lowered by natural surfactants (such as proteins, two or more of the above behaviors.
lipoproteins, polypeptides, and fatty acids) to about 45-50 dyn/ The surface-active agents (e.g., proteins) that cause foam
cm; addition of surfactants can lower surface tensions substan- emerge from media components, extracellular enzymes, or cell
tially to values as low as 28 dyn/cm (Evans and Hall, 1971; lysis (van’t Riet and Tramper, 1991). The presence of proteins
Hall, 1971). Foams also can be stabilized by these surfactants causes foaminess at concentrations from 1 mg/L up to 1-10
(Jenkins et al., 1993). g/L, after which further concentration increases decrease foami-
Conditions That Affect Foaming. Some amount of foaming ness since proteins coagulate (van’t Riet and Tramper, 1991).
during fermentation is acceptable, but excessive foaming For example, for Penicillium cultivations in the absence of
requires some type of control action (van der Pol et al., 1983; antifoam, foaminess has been found to increase, pass through a
Prins and Van’t Riet, 1987). Conditions that affect the degree maximum, and then decrease (König et al., 1979).
of foaming during fermentation include gas introduction (i.e., Medium Composition and the Presence of Cells, Particles,
aeration), medium composition, cell growth, metabolite forma- and Surfactants. Many aspects of the initial medium composi-
tion, surface-active substance formation, and indirectly, vessel tion affect foam formation, including concentrations of salts,
geometry (Taticek et al., 1991; Vardar-Sukan, 1992). Knowledge proteins, and sugars, as well as the presence of alcohols (Vardar-
of these conditions permits optimization of fermentation pro- Sukan, 1992). For example, for soybean-based media formula-
cesses to minimize factors that may cause foam such as oxygen tions, the acid-hydrolyzed product generated more foam, owing
starvation and nitrogen limitation (Nobel et al., 1994b). Neural to its higher oil content, than the microbe-hydrolyzed product
networks have been used to predict foam behavior and recom- (Hall, 1971). Increases in glucose up to 24 wt % in medium
mend specific action (i.e., add/not add antifoam, adjust process containing 1-3% soybean meal increased foam stability,
conditions) (Brown et al., 2001). possibly by raising medium viscosity (Hall, 1971). Thus, the
Quantification. Foaminess is a characteristic of the solution selection of medium components, including their method of
having units of time, related to the mean residence time of gas manufacture and their concentration, can impact foaming. In
in foam (Lee et al., 1993). Foaminess is defined as the fact, the protein content of growth medium has been cited as
equilibrium volume of foam/volumetric flowrate of gas or the the most vulnerable factor for foaming (Vidyarthi et al., 2000).
height of foam/superficial velocity for cylindrical tanks (Bum- Salt-forming substances generally increase foaming and
bullis et al., 1979; Lee et al., 1993). Quantification of foam surface viscosity as well as decrease surface tension (Ghildyal
volume and sometimes even foam height is difficult at the large et al., 1988). Although solution foaminess increases with
scale (Bumbullis et al., 1979). Foaminess is a measure of increasing salt concentration, foam stability diminishes, most
foaming capacity, independent of equipment geometry and likely owing to changes in aqueous protein stability due to salt
measurement techniques but dependent on media ingredients, concentration (Vardar-Sukan, 1992; Kotsaridu et al., 1983a;
their relative concentrations, and various physical factors Bumbullis et al., 1979). Specifically, protein solubility increases
(Ghildyal et al., 1988). Raising gas flowrates to increase foaming at lower salt concentrations (“salting in” effect), which in turn
for solutions with low foaminess is hampered by difficulty increases foam stability. Protein solubility decreases at higher
detecting the foam/liquid interface when high air flowrates are salt concentrations, raising apparent protein concentrations in
present (Edwards et al., 1982). Other relevant measures to the foam layer, which in turn increases foaminess (Prins and
characterize foaming include the liquid volume held in the foam van’t Riet, 1987). Interestingly, low molecular weight (3000-
(foam volume/liquid volume before foaming), volumetric foam 10,000 Da) proteins and glycoproteins promoted foam more than
overflow rate, and foam volume decrease over time (Wong- higher molecular weight ones (>50,000 Da) (Hall, 1971).
samuth and Doran, 1994; Abdullah et al., 2000). Short chain alcohols (such as methanol, ethanol, or propanol)
Fermentation Phase. The formation of foam differs depend- increase foaminess of protein/water solutions with a maximum
ing on the fermentation phase. The two different foam phases effect at 1-2% v/v (van’t Riet and Tramper, 1991; Vardar-
are (1) early in the fermentation, for example, owing to proteins Sukan, 1992). Alcohol additives also increase effective protein
Biotechnol. Prog., 2007, Vol. 23, No. 4 769

Table 1. Summary of the Influence of Fermentation Operating Conditions on Foam Formation.


element general influence mechanism comments reference
gas flow increases with higher greater amount of bubbles often foam maximum Vardar-Sukan, 1992;
rate flow rates erupting from surface exists and then decreases Van’t Riet and Tramper, 1991
superficial increases with higher greater speed of bubbles if vvm held constant with Pandit, 1989
velocity velocities erupting from surface scale up, superficial velocity
increases
sparger increases with smaller smaller bubbles form sintered spargers may Chisti, 1993
orifice orifices smaller foam cell produce high amounts
size structures which of foam
collapse more slowly
agitation increases with higher gas entrainment; cell depending on relative Hoeks et al., 1997,2003;
rate rates lysis owing to high broth/impeller geometry, Pandit, 1989
shear environment possibly can decrease foam
by providing mechanical
disturbance
viscosity increases as viscosity decreases film drainage system-dependent behavior Prins and Van’t Riet, 1987
rises above initial which increases foam
water-like values
temper- decreases with higher decreases viscosity which foaming can increase when Prins and Van’t Riet, 1987;
ature temperature increases liquid film broth cooled awaiting Gaden and Kevorkian, 1956
drainage and reduces harvest
foam
broth increases when near proteins least soluble foaming can increase Vardar-Sukan, 1992;
pH protein isoelectric near pI protein denaturation van’t Riet and Tramper, 1991
point (pI)
sterili- increases with longer Maillard reaction products foam may decrease after Kotsaridu et al., 1983b;
zation sterilization hold formed from sterilizing aeration applied for a Vardar-Sukan, 1992;
times, higher temper- nitrogen sources and duration since Maillard Schügerl, 1985
ature, higher pre-steri- reducing sugars together reactions partially
lization pH reversible

concentrations in the foam layer, causing higher foaminess but For example, stable foams are formed when the yeast Moniliella
reducing stability in some instances (Bumbullis and Schügerl, became nitrogen-limited and are stabilized by secreted polysac-
1979; Wilde et al., 2003). In addition, carbohydrates and R-keto charides. This foam could not be suppressed using antifoam or
acids enhance protein-containing foam stability, presumably by mechanically destroyed using foam breakers (Burschapers et
affecting gas-liquid interfacial activities (Noble et al., 1994b). al., 2002). Removal of cells from medium due to foam formation
Components of distillery fermentation broths, which include can cause autolysis, which releases microbial proteins that
sugars, electrolytes, proteins, and acids, concentrate at the gas- enhance foam stability (Stanbury and Whitaker, 1984). Micro-
liquid interface causing elastic foam that entraps fermentation organisms trapped in foam experience oxygen and nutrient
gases (Romualdo et al., 2002). Lipids, consisting of longer chain limitations (Pandit, 1989; Vogel, 1983), causing changes in
fatty acids (g12 carbons), reduce foam in beer, competing with microbial metabolism, protein denaturation, and/or microbial
proteins to weaken liquid films and destabilize bubbles (Wilde lysis (Vardar-Sukan, 1992). Proteins and microorganisms also
et al., 2003). are concentrated by froth flotation (Vardar-Sukan, 1992).
The presence of cells, as well as secreted compounds, can Cells caught in stable foams form crusts or meringues, often
influence foam formation and stability. The presence of finely
adhering to walls (i.e., forming a ring) and cause sampling non-
divided, insoluble particles can assist in foam stabilization
uniformity (Abdullah et al., 2000). This problem is especially
(Schügerl, 1985) by concentrating their presence at gas-liquid
prevalent in plant (Wongsamuth and Doran, 1994) and certain
interfaces (Gaden and Kevorkian, 1956). Specifically, aqueous
fungal cultures. Many plant cells secrete polysaccharides and/
solutions of colloidal materials adsorb at surfaces (i.e., inter-
or proteins, causing cultures to become sticky, entrapping
faces) and add mechanical strength to the films formed (Gaden
bubbles in foam that then entraps cells forming a crust (Taticek
and Kevorkian, 1956). In fermentation applications, foams can
et al., 1991).
be stabilized by hydrophobic particles, such as Nocardia cell
solids, since this culture has long-chain, hydrophobic mycolic The onset of substantial cell growth reduces foaming,
acids on its cell surface (Jenkins et al., 1993). Generally, thick suggesting metabolism modifies the composition of surface-
cell suspensions exhibit less foaming than dilute broths perhaps active agents (Bungay et al., 1960); thus some metabolites
as a result of higher viscosity, but at very high viscosities foam directly and others indirectly act as antifoams (Soifer et al.,
can reappear (van’t Riet and Tramper, 1991). In some thick 1974). Foaming decreases in extent and duration with greater
mycelial fermentations, there is only a gradual increase in the inoculum age during production fermentations of Actinomyces
gas proportion in liquid from the bottom up to the top of the streptomycini owing to increases in antifoaming metabolite
liquid height, making it harder to quantify foam levels. However, concentrations (Soifer et al., 1974). In fact, the bacteria
the effect of cells is hard to distinguish since cells are always Xenorhabdus has been demonstrated to produce antifoam during
present together with some amount of protein (van’t Riet and its normal metabolism (Jewell and Dunphy, 1996). The creation
Tramper, 1991). (via recombination by crossing strains) and/or selection (via
Foams in fermentations are likely derived from a variety of mutation) of a non-foaming strain of a commercial organism
excreted products or cell lysis products and not solely from helped control foaming late in fermentation (Stanbury and
extracellular proteins (Vardar-Sukan, 1992; Noble et al., 1994ab). Whitaker, 1984). The use of such foam-negative mutants
770 Biotechnol. Prog., 2007, Vol. 23, No. 4

(Ishizuka et al., 1989) avoids changes in fermenter operating and van’t Riet, 1987). Liquid-phase surface tension directly
conditions and reduces antifoam additions. affects stable foam height (Pandit, 1989). The liquid film
Fermentation Operating Conditions. Fermentation operating thickness forming the foam bubble boundary is proportional to
conditions strongly impact on the initiation and severity of foam σ0.57 (Pandit, 1989). Ionic strength primarily affects foaming
formation. High air flowrates, coupled with foam-stabilizing by altering protein solubility as discussed previously.
proteins and carbohydrates present in the broth, make fermenta- Broth pH affects the action of antifoam agents (Vardar-Sukan,
tion processes prone to foaming and particularly challenging 1992; van’t Riet and Tramper, 1991) because it affects foams
applications for antifoams (Pelton, 2002). A summary of these produced by colloidal agents such as proteins (Gaden and
trends is shown in Table 1. Kevorkian, 1956). If the pH is near the protein pI (isoelectric
Foam levels generally increase in height with increasing gas point) where proteins are least soluble (Vardar-Sukan, 1992),
flow rate since more bubbles erupt from the liquid surface and foam generally reaches its maximum extent (van’t Riet and
then are converted into foam (van’t Riet and Tramper, 1991; Tramper, 1991). In one example, broth pH decreases from 6 to
Vardar-Sukan, 1992). In many cases, a maximum occurs after 4 increased foaming and foam stability about 10-fold for a
which foam layers decrease in size as gas flowrates become medium containing 4-5% soybean meal (Hall, 1971). In a
higher (König et al., 1979). This behavior is possibly due to a second example, foam volume increased 4-fold and foam
decrease in bubble monodispersity since higher flowrates raise stability increased 6-fold for shake flask plant cell cultures of
coalescence rates (van’t Riet and Tramper, 1991) or an increase Atropa when broth pH was lowered from 7 to 6 (Wongsamuth
in mechanical disturbances when higher flowrates disengage and Doran, 1994).
from the broth (Prins and van’t Riet, 1987). Sterilization increases the already high foaming capacity of
The stable foam height generally increases directly with complex nutrient media considerably (Vardar-Sukan, 1992;
higher gas velocities and greater liquid depth above the sparger Schügerl, 1985). Heat causes nitrogen sources to become
(Pandit, 1989). Gas superficial velocities themselves increase hydrolyzed or partially degraded, leading to Maillard reactions
upon scale-up if specific aeration rates (i.e., volume air per between reducing sugars and amino acids/proteins/peptides;
volume broth per unit time) remain constant. Consequently, these Maillard reaction products enhance foam formation
superficial velocities relevant to the large scale (e.g., 600 L) (Vardar-Sukan, 1992). Maillard reaction products increase with
should be examined during smaller scale (e.g., 20 L) bioreactor higher sterilization temperatures, longer sterilization times, and
process development work (Hoeks et al., 1997). higher presterilization pH values (Kotsaridu et al., 1983b).
Sparger orifice designs impact foaming by affecting bubble Overall Maillard reaction products can increase foaminess by a
size. For hybridoma cells cultivated in serum-containing media, factor of 2100 during sterilization of potato-protein liquor and
porous metal spargers (180-200 µm, 0.00018-0.0002 m) glucose medium (Ghildyal et al., 1988). During sterilization
produce foams with bubble sizes of about 0.002-0.003 m. These foam-stabilizing components form in soybean flour medium
foams are challenging to control because they are densely based on Maillard reaction product formation (Vardar-Sukan,
packed. In contrast, foams produced by ring spargers, with holes 1992). Since the Maillard reaction is partially reversible, a
<0.001 m that emit bubbles of about 0.01-0.02 m, are easier reduction of foaminess is observed when sterilized medium is
to control because the larger bubbles formed large foam cell “aged” by aeration with sterile air (Vardar-Sukan, 1992) or
structures and most bubbles collapsed quickly (Chisti, 1993). stored at low temperatures (Kotsaridu et al., 1983b). Other
Larger holes in sparger rings also can reduce foaming in plant specific examples of these effects have been reported: (1) The
cultures (Taticek et al., 1991). foam stability of a medium containing 2% soybean meal, 4%
Agitation often increases foam by increasing air entrapment glucose, and 0.5% CaCO3 increases 5-fold during a 90 min
and cell lysis. As impeller speed increases, foam cell size sterilization at 125 °C (Hall, 1971). (2) The foaming coefficient
decreases and becomes more stable, which in turn increases the of molasses increases 2-fold for a sterilization temperature
rate of foam buildup (Pandit, 1989). Once foam has formed, increase from 110 to 130 °C (Berovič, 1992).
however, increased agitation sometimes reduces foam height Consequences of Foam. Sterilization. Foam during steriliza-
owing to mechanical disturbance. However, the effectiveness tion can cause sterility problems during the subsequent cultiva-
of this measure depends highly on impeller position relative to tion (Vardar-Sukan, 1992). Foaming during sterilization splashes
broth level. There is additional anecdotal evidence for animal media (e.g., flour) particles on fermenter side walls that may
cell cultures that downward pumping hydrofoil impellers reduce be sterilized inadequately. Foam probes can be enabled during
foam formation relative to other impeller designs such as sterilization to detect foam formation.
Rushton impellers. Working Volume. Foaming enhances gas holdup (Vardar-
Foam level and persistence decreases as temperature increases Sukan, 1992), which expands broth volume (Berovič, 1992),
potentially as a result of decreased broth viscosity (Vardar- often requiring reductions in the working (i.e., operating) volume
Sukan, 1992; Prins and van’t Riet, 1987), increased drainage of the fermenter below the maximum level to ensure sufficient
of liquid films (van’t Riet and Tramper, 1991; Gaden and headspace (Edwards et al., 1982; Brown et al., 2001). These
Kevorkian, 1956), and lower gas pressure within bubbles (Gaden effective liquid volume increases lead to broth (e.g., substrate,
and Kevorkian, 1956). These observations possibly explain biomass) loss through air entrainment, creating several sterility
foaming increases when broth is cooled while awaiting harvest. as well as containment problems (Bryant, 1970; Vardar-Sukan,
Other authors found that temperature increases decrease foam 1992; Berovič, 1992). Wetting of off-gas hydrophobic vent
stability but enhance foaminess (Ghildyal et al., 1988) since filters (Duitschaever et al., 1988) causes plugging and back-
protein denaturation increases (Vardar-Sukan, 1992). pressure buildup and then potential release of pressure relief
Foam formation is affected by the liquid (broth) properties, devices (Koller, 2004). Cells that become entrapped in foam
specifically viscosity, surface tension, and ionic strength (Var- via flotation (Bryant, 1970) or deposited on upper parts of the
dar-Sukan, 1992). Foam increases when viscosity rises from 1 fermenter undergo autolysis, which releases surface-active
mPa-S up to 10-100 mPa-S; beyond 10-100 mPa-S foaming agents that catalyze further foam formation (Ghildyal et al.,
decreases, but at very high viscosities foaming starts again (Prins 1988; Vardar-Sukan, 1992). Productivity also is reduced from
Biotechnol. Prog., 2007, Vol. 23, No. 4 771

foam-related depositions of substrate and/or product on non- high-frequency signals with wide band measuring bridges and
wetted parts of the vessel and piping (Brown et al., 2001; Pandit, amplifiers able to measure foam even with a biofilm present
1989). (PharmaTec, 2004). Another alternative for foam detection is
Mass Transfer. Mass transfer is enhanced when sur- sonar level detection based on the presence of a foam layer in
face-active compounds improve bubble stability by reducing some cases interfering with sonar feedback.
coalescence and increasing the surface area available for mass Control Strategies. Several early examples of automated foam
transfer (Noble et al., 1994b). However, two mechanisms exist detection and antifoam addition systems are noted in the
for antifoams to reduce mass transfer: (1) creation of larger literature (Pfeifer and Heger, 1957; Bartholomew and Koslow,
bubbles and decreased gas holdup due to antifoam-induced 1957), forming the basis for how antifoam addition is controlled
bubble coalescence, and (2) creation of a spread film by the today. Since on-off foam detection only indicates that foam is
defoamer on bubble surfaces, reducing oxygen diffusion rates present (Brown et al., 2001), it is being replaced by continuous
(Pelton, 2002). In the latter mechanism, mass transfer is slowed level detection that gives the foam height and permits set point
due to antifoam accumulation at the gas-liquid interface causing adjustment. For sensors linked to automated antifoam addition
an additional gas-liquid interfacial resistance and decreased systems, small time lags are introduced to prevent over-dosing
gas diffusion (Vardar-Sukan, 1992). Mass transfer coefficients and control the stop/start operation of the pump (PharmaTec,
are relatively independent of DO (i.e., concentration driving 2004). The time lags also permit adjustment of (1) the amount
force) without antifoam present; with antifoam present these of antifoam added, (2) the interval between antifoam additions
mass transfer coefficients become linearly dependent on DO, to permit mixing and for the antifoam to take effect, and (3)
possibly due to the oxygen diffusion coefficient through the the detection sensitivity threshold by requiring continuous
antifoam layer becoming concentration-dependent, an effect detection for a designated time period to avoid additions caused
noted for polymers (Bull and Kempe, 1971). by splashing (Reisman, 1988; Getchell, 1983). In one example
Foam not only causes changes in air bubble size but also in strategy, antifoam addition begins if the foam signal persists
composition, altering dissolved gas profiles due to heterogeneous for 5 s continuously, then alternately adds antifoam for 30 s
gas dispersion (Vardar-Sukan, 1992). Foam increases bubble and mixes without antifoam being added for 30 s, repeating
residence time in broth, resulting in bubbles becoming oxygen- this cycle until foam is not detected for 5 s before stopping.
depleted (Stanbury et al., 1995) and accumulating carbon For early antifoam addition on/off cycles, incorporation of a
dioxide, thus delaying its removal (Berovič, 1992; Koch et al., time lag permitted mixing and reduced overall antifoam
1995). Interestingly during early hybridoma animal cell cultiva- consumption when compared to manual addition methods
tions at the 1 L scale, stable foams are observed to increase (Bartholomew and Koslow, 1957; Dworschack et al., 1954).
surface aeration by 90% (Ju and Armiger, 1990), but this effect The antifoam addition tubing bore size and pump rate can be
is considered primarily an indirect result of reduced survival of used to estimate the amount of antifoam added per shot
cells present in foams. (although this value is somewhat dependent on fermenter back
Power Input/Mixing. The presence of foam reduces apparent pressure), or antifoam reservoirs can be weighed continuously.
viscosity owing to higher gas holdup, thus decreasing power Prevention of Foam without Adding Antifoam. Foam
dissipation and circulation rates (Vardar-Sukan, 1992). Thus, prevention, without additional antifoam beyond what may be
permitting some foaming (and avoiding antifoam addition) can present in the initial medium composition, is preferred if it is
increase mass transfer substantially and reduce power consump- reliably attainable with no deleterious effects on the process
tion (Berovič, 1992). In one example, when medium is changed (Vardar-Sukan, 1992). A brief summary of available options is
from a non-foaming to foaming system, a higher OTR is presented below.
achieved at same power input (Adler and Fiechter, 1986). Physical and Mechanical Methods Including Foam Break-
Instrumentation. Foam on the broth surface adversely affects ers. Physical methods to prevent foam such as ultrasound,
level measurements, which makes continuous cultivation dif- thermal, or electrical treatment can adversely affect cells
ficult (Berovič, 1992). The presence of small bubbles from foam (Vardar-Sukan, 1992) so these are not discussed further.
also can interfere with electrode sensors (Vardar-Sukan, 1992; Mechanical methods reduce foam by subjecting it to shear
Pandit, 1989). stress, but these devices can have significant extra power
Miscellaneous. In some difficult fermentations (no specific requirements (Brown et al., 2001). In some instances, they can
information given by authors), foam must be strictly avoided increase cell death despite similar biomass concentrations (Vraná
because certain broth components are rendered corrosive by high and Seichert, 1988), and they may preferentially destroy larger
oxygen levels in foam (PharmaTec, 2004). over smaller foam bubbles. Common mechanical foam breakers
Detection of Foam. Types of Foam Sensors. Several types use rotating elements such as discs, bladed wheels, or stirrers
of foam sensors exist to detect the presence and, in some cases, (Solomons, 1969; PharmaTec, 2004; Yasukawa et al., 1991;
depth of foam. Conductive probes, based on conduction of Takesono et al., 2001) or a simple bar attached to the agitator
electrical charge (i.e., DC voltage) between the probe and vessel shaft above the liquid level. They often are complicated in
by ions present in the liquid (Brown et al., 2001), are sensitive design, have high running costs, and also are unreliable (Prins
to biofilm (Hall, 1971), which causes a permanent short circuit and van’t Riet, 1987; Vardar-Sukan, 1992). Mechanical foam
due to the collection of material across the probe’s insulation breakers are used when the process cannot tolerate chemical
(Ghildyal et al., 1988). Capacitance probes have less biofilm antifoams (Olivieri et al., 1993), as was the case for early animal
interference than conductive probes since the entire capacitance cell cultivation processes (Chisti, 1992). Foam breakers, used
probe is covered with insulation (Hall, 1971). Capacitance is in conjunction with antifoam addition, can reduce antifoam
sensed as the foam becomes dielectric (Ghildyal et al., 1988). addition by 33-50% (system not given) (Yamashita, 1972).
Both conductance and capacitance foam sensors are able to Agitation Rate. Although counterintuitive, stirring-as-foam-
measure liquid (i.e., foam) level (Getchell, 1983; Brown et al., disruption (SAFD) reduces foam by causing a higher liquid
2001), possessing a user-adjustable detection level rather than velocity to be generated by an upper impeller strategically
fixed point. Less common are admittance probes, which use located near the broth-foam (dispersion) interface. Axial
772 Biotechnol. Prog., 2007, Vol. 23, No. 4

impellers are more effective at this task than Rushton impellers. for penicillin fermentation, 50-70% decrease in mass transfer
Specifically, for the volume range where this technique is (Chain et al., 1966). The amount also varies depending on the
effective, the foam height is almost linearly dependent on broth antifoam class for fermentation of the yeast Torulopsis on
mass for constant Vs and rpm (Boon et al., 2000; Hoeks et al., complex medium containing sugar beet molasses: control 47;
1997,2003). The mechanism of SAFD is foam entrainment; the 0.1% Dow Corning Antifoam A (Dow-Corning, silicone-based)
gas holdup of the broth rises sharply when the impeller knocks 18.8; 0.1% Hodag K-4 (Lambent, organic-based) 20.4; 0.1%
down foam. Thus, the aerated liquid volume relative to the upper lard oil (natural oil) 31.4, all in units of mmol/L-h (Phillips et
impeller location becomes a key parameter for effectiveness al., 1960). Finally, the amount varies based on antifoam quantity
(Boon et al., 2002). In this pilot plant facility, this technique is present: specifically, the KLa of pure medium decreases sharply
effective for a fungal Glarea fermentation, as well as more upon addition of a small amount of KM70 (silicone emulsion)
effective than P2000 antifoam addition for knocking down foam antifoam (Shinetsu Kagaku); after about 200 ppm is added there
induced by addition of Tween 80 surfactant. is no further reduction (Takahashi and Yoshida, 1979). Similarly,
Other Operating Conditions. Foam generally is reduced by mass transfer rates decrease from 0 to 0.01% antifoam with no
increasing back-pressure, decreasing agitation, and/or decreasing effect between 0.01% and 0.1% for Desmophen 3600 (polyester
aeration during cultivation. Reducing airflow or agitation rates polyol) added to protein solutions (Adler et al., 1980). A critical
to reduce foam usually adversely affects productivity (Vardar- antifoam concentration does exist above which both mass
Sukan, 1992). Although productivity often is less affected by transfer and gas holdup increase (Morão et al., 1999; Kawase
back-pressure, higher back-pressures may not always be attain- and Moo-Young, 1990). Owing to greater gas holdup and lower
able at production scales without limiting airflow rates. Increas- circulation rates, foaming also can result in lower heat as well
ing back-pressure from 10-15 to 30 psig decreases foam during as mass transfer rates (Vardar-Sukan, 1988).
production of itaconic acid by Aspergillus (Pfeifer et al., 1952). When antifoam effects are measured in water or media
Increasing back-pressure from 1.1 to 1.5 kg/cm2 when applying without cells, there are steep decreases in KLa reported, most
air to end sterilization and until the batch is cooled to its likely due to most of the antifoam accumulating around the gas-
cultivation temperature also served to reduce foam in this pilot liquid interface and interfering with mass transfer. In one
plant facility. example using 3000 L fermenters, the mass transfer decrease
Impact of Antifoams. Several negative effects of antifoams caused by antifoams is largest (70-85% decrease) in sulfite
are noted in the literature. In general, negative effects of solutions, less (50-70% decrease) in P. chrysogenum, and even
antifoams decrease when they are added regularly to fermenta- lower (25% decrease) in the yeast Torula utilis (Chain et al.,
tion processes in low amounts rather than in fewer additions of 1966). These results suggest that to truly characterize the effects
higher amounts (Kovalev et al., 1982). of antifoam on mass transfer, detrimental effects should be
Concentration. Foam formation can be enhanced by anti- characterized comparing antifoam-containing and antifoam-free
foams present in too high concentrations (Vardar-Sukan, 1992). media containing cells (i.e., actual fermentation broth) since
Specifically, foam stability decreases up to a maximum surfac- antifoam distribution is affected by the presence of hydrophobic
tant concentration, and then subsequent surfactant additions cells.
actually increase foam stability (Vardar-Sukan, 1992). When Power Input/Mixing. Depending on the antifoam concentra-
too much antifoam is added, foam is destroyed, but the liquid tion, sharp increases in power draw due to degassing or a drop
becomes strongly coalescing (Prins and van’t Riet, 1987). in mixing power due to lowered gas entrainment is observed
Mass Transfer. A tradeoff exists between minimizing foam (Ghildyal et al., 1988). In one instance, an increased power draw
height and maximizing mass transfer (van’t Riet and van of up to 20% for large fermenters is observed after addition of
Sonsbeek, 1992). Antifoams present at low concentrations a natural oil antifoam (Bungay et al., 1960).
markedly decrease the volumetric mass transfer coefficient, KLa, Biological and Metabolic. The presence of antifoam can
with the major effect being on the overall gas-liquid mass change the growth rate and even morphology of cells (Noble
transfer coefficient, KL, and some effect on the specific et al., 1994b; PharmaTec, 2004), also negatively affecting cell
interfacial area, a (Morão et al., 1999; Kawase and Moo-Young, density and/or product concentration (PharmaTec, 2004). Certain
1990). Conditions that cause bubble collapse in foam also antifoams are preferentially used as a carbon source by some
promote coalescence of bubbles within the liquid phase, which cultures (Ghildyal et al., 1988). The broth pH profile can be
results in larger bubbles with lower specific interfacial surface altered by fatty acids released due to lipase action on oils which
areas and thus lower mass transfer (van’t Riet and van Sonsbeek, both changes culture metabolism and fermentation progression
1992). Different types and quantities of antifoams can affect (Ghildyal et al., 1988; Elander, 1989). In one application, this
dissolved gas dispersion to the culture and consequently overall behavior is used to cause crude pH adjustment by relying on
respiration rate (Elander, 1989). When gas holdup and specific the culture to produce fatty acids by metabolizing antifoam oils
interfacial area are reduced suddenly by anti-foam addition, the (Bungay et al., 1960).
broth DO drops quickly (Schügerl, 1988; Yagi and Yoshida, Instrumentation. In one report, antifoams are found to
1974; Lengyel and Nyiri, 1966) but also recovers quickly. accumulate on the membrane surfaces of DO probes, increasing
In general, mass transfer rates can be lowered by antifoam diffusion resistance (Chen and Wang, 1993). However, this
up to 50% (Berovič and Cimerman, 1979; Solomons, 1966; study is conducted in the absence of cells; in the presence of
Atkinson and Mavituna, 1983) or 60% (PharmaTec, 2004) or cells, the cells themselves also are observed to cause significant
75% (Phillips et al., 1960). For example, despite higher resistance. There are other informal reports of antifoam ac-
interfacial area owing to a 13% reduction in surface tension, cumulation when concentrations exceed 1 mL/L on the surface
0.25% of 3% Alkaterge C in lard oil causes a 50% reduction in of filter probes, designed to permit cell-free sampling of
the gas (oxygen) absorption coefficient of filtered Penicillium fermentation broths.
broth using complex medium (Deindoerfer and Gaden, 1955). Downstream Isolation. Antifoams retain their surface-active
The amount of reduction varies depending on the fermentation nature during isolation steps, so their permissible levels during
class: for yeast fermentation, 25% decrease in mass transfer; mid-cycle fermentation should be examined in conjunction with
Biotechnol. Prog., 2007, Vol. 23, No. 4 773

downstream processing requirements (Paul et al., 1981). Anti- Table 2. Selected Silicone-Based Oil (S) and Emulsion (SE)
foams cause difficulty in extraction and subsequent purification Antifoams Noted in Fermentation Literaturea
owing to the formation of difficult-to-break emulsions in past and/or current
aqueous-solvent systems (Ghildyal et al., 1988), specifically tradename vendor
impairing phase separation during whole broth extraction (Paul Silicone (100%) [Poly(methylsiloxane) in Silicone Oil]
et al., 1981). Antifoams also can co-extract with the desired Sag M-10 Dow Corning
product (Pollard et al., 2006). Antifoams accumulate on mi- Sag 471 Union Carbide
DC-A, A Dow Corning
crofiltration membranes causing membrane fouling and even Antifoam A (concentrate) Sigma
destruction (PharmaTec, 2004; Liew et al., 1997). Often Mazu DF100, DF1005 Noveon, Mazur
changing the filtration step temperature relative to the antifoam’s FD 101b Stepan
cloud point can enhance performance, however, and an in- Q7-2243 LVAb Dow Corning
C100F b Basildon
process assay to detect their removal is critical to efficiently S184 Wacker Chemie
developing these steps. Finally, antifoams interfere with the Q10-335b Dow Corning
binding step in ion exchange separations. Because of these TMA812 Toshiba
isolation interferences, extra separation steps, often specific in Antaphron NM40 UK
Biospumex FDA 165K Cognis
nature to the antifoam employed, are implemented to remove
defoamers (Ghildyal et al., 1988). Silicone Emulsions (Silicone Percentage)
KM-70, M-7W-0 (37%) Shinetsu Kogaku
Mazu DF 210, 210 S, 2105 (10%) Mazur/Basf/Noveon
Chemical Antifoams (Defoamers) 1510 (10%) Dow Corning
1520 (30%) Dow Corning
Antifoams and defoamers are fundamentally the same chemi- AF (30%) Dow Corning
cals despite some differing implications noted in the literature C (30%) Dow Corning
(Pelton, 2002). Antifoams are defined as strongly surface-active DC-B (10%) Dow Corning
substances which replace foam-forming components and lower FG-10 (10%) Dow Corning
DSP (10%) Dow Corning
surface tension of liquids (van’t Riet and Tramper, 1991; van’t RD (10%) Dow Corning
Riet and van Sonsbeek, 1992). Antifoams are dispersed by M30 (30%) b Dow Corning
stirring, and foam is destroyed by bubble coalescence, which Q7-2587 (30%) Dow Corning
decreases the available surface area for gas-liquid mass transfer Hodag FD-62 (10%) Lambent
(de Haut, 2001). Antifoams typically are added to medium or Hodag FD-82 K (30%) Lambent
A (A-5758) (30%) Sigma
broth before foaming occurs (Ghildyal et al., 1988), and B (10%) Sigma, JT Baker
sometimes they are called foam inhibitors (Weng et al., 1997). C (30%) Sigma
In contrast, defoamers compete with other surface-active agents Chemax DF-10 (10%) Rutgers
for the surface layer, but they do not support foam formation Chemax DF-30 (30%) Rutgers
SE15 (10%) Sigma
(Gaden and Kevorkian, 1956). Defoamers are self-dispersed, SE9 (10%) Wacker Chemie
and foam is destroyed by surface action (de Haut, 2001). Sag 10 (10%) Union Carbide
Typically, defoamers are used to knock down foam after it has Sag 30 (30%) Union Carbide
formed (Ghildyal et al., 1988), and sometimes they are called Silcolapse 5001 (10%) ACC/Rhodia
Silcolapse 5000 (30%) ICI/Ambersil/Rhodia
foam breakers (Weng et al., 1997). For the purpose of this
GE 60 (30%) GE
review, the term antifoam is used to cover both aspects. Roth 0865 (UK%) Carlroth.de
Antifoams act via four steps via antifoam droplets: (1) PD30 (30%)b Basildon
entering the liquid film around the foam bubble, (2) bridging RS-70426 (27%)b Rhodia
the width of the liquid film, (3) dewetting by causing the liquid Silicon-Based
film to thin around where the antifoam is present, and (4) rupture 1705-W Lion
Assaff III Rhone-Poulenc
of the liquid film (Garrett, 1993; Denkov et al., 1999; Jha et
al., 2000; Christiano and Fey, 2003). Certain antifoams contain Complex Polyhydric Alcohol/Silicone Polymer
SO-25b Sigma
insoluble particles which reduce the viscosity of foam lamellae,
a Information obtained from company websites, including MSDSs, and
causing bubbles to thin and drain in this fashion, and severely
reducing foam stability (PharmaTec, 2004). handbooks (Ash and Ash, 2000a,b) in addition to publications. Information
limited by proprietary disclosures. Silicone oil is a silicone polymer, also
Chemical antifoams are simple and economical (Vardar- known as poly(dimethylsiloxane). Simethicone consists of 95% poly(di-
Sukan, 1992). Consequently, they have proliferated and an methylsiloxane) (silicone oil) and 5% hydrophobic silica. Sigma Antifoam
abundance of choices are commercially available, both currently A/C emulsions are both 30% but use different emulsifiers. Sigma sells both
Antifoam A concentrate (100% silicone) and emulsion (30%). Some authors
and in the past. Tables 2, 3, and 4 compile examples of do not distinguish which type was used. b Antifoams offered commercially
antifoams, many of which are cited in fermentation literature. but not cited in fermentation literature at this time.
Unfortunately, often trade names do not aid in understanding
the nature of the compounds. Consequently, similar antifoams
Desired Characteristics. No single antifoam can possess all
are grouped together by “class” within Tables 2, 3, and 4 so
of the ideal, desirable characteristics (Vardar-Sukan, 1992), thus
that available antifoams options for fermentation applications
are readily accessible and assessable. Confidentiality concerns a compromise often is needed. Antifoam efficiency is directly
frequently preclude obtaining precise details about antifoam proportional to its foam suppression ability and inversely
compositions. In a few cases, little or no information is proportional to the amount consumed in the fermentation to
obtainable. For those cases where an antifoam is composed of suppress foam (Vardar-Sukan, 1988). This efficiency is deter-
components from more than one antifoam class, it is listed with mined by comparing the (1) minimum volume required and (2)
the class of its highest percentage composition. In instances maximum yield of product or absence of any microbial activity
where one antifoam class is dissolved in a carrier of a second (Duitschaever, 1988). Overall antifoams must have properties
antifoam class, the class of the solute is applied. that permit them to function as antifoams, be used with living
774 Biotechnol. Prog., 2007, Vol. 23, No. 4

Table 3. Selected Poly(alkylene glycol) (PAG)-Based Antifoams Noted in Fermentation Literature


chemical tradename vendor
poly(propylene glycol) (linear polyether of propylene oxide), PPG P2000 Dow
PPG2000 Wacker-Chemie
PPG 2025 E. Merck
PPG1025 (mw 1000) BDH
PPG-24 butyl ethera UK
Pluriol P2000a BASF
Pluracol P2000a BASF
Pluracol P2010 BASF
Emkaphyl Dow
poly(ethylene glycol) (linear polyether of ethylene oxide), PEG Pluracol E BASF
PEG Typ 300 Carl Roth
poly(alkylene glycol) (e.g., linear polyether of ethylene and KFO F119, F161a Noveon
propylene oxides; poly(propylene glycol) KFO 673 KABO/Lubrizol/Emerald
ether of butyl alcohol) UCON LB625 (mw 1500) Dow
Mazu DF60P Mazur
Mazu DF7960 Mazur
Mazu DF800 S, DF8005 Mazur
Ucolub N115, N-7 or N-7/1 Brenntag
Struktol J647, SB2121 Schill and Seilacher
SP1 Th. Goldschmidt
difunctional ethylene/propylene oxide (EO/PO) block copolymers Tetronic 901 BASF
Pluronic F68 BASF
Pluronic L 61, PL-61, PE6100 Ugine Kuhlmann
Pluriol L81, PE8100 BASF
Pluronic L122 BASF
Pleuronic Fluka
Adekanol LG-109, LG-107 Fermenta AB, Asahi Denka Kogyo, Adeka
Mazu DF204 PPG Ouvrie
polyalcohol based on EO/PO block copolymer Struktol J650a Schill and Seilacher
poly(alkylene glycol) (PAG)-based Breox FMT30 Intl. Specialty Chemicals/Cognis
polyether Tego KS911 Degussa, Goldschmidt
likely PAG-based: 99% organic defoamer, 1% silicon glycol to Antifoam 289, 286 Sigma
enhance spreadability
PAG-based: 20% branched simethylsiloxane/3.5% dispersed VP1133 Wacker-Chemie
silicone dioxide/76.5% PPG2000
PPG-based polyether dispersions Antifoam 204 Sigma
PAG-based with small amount of silicone XFO-371 Ivanhoe
PAG-based: 95% polyalkylene/5% silicone Sag 5693, 5698 Union Carbide, OSI Specialties
PAG derivative Disfoam GD Nihon Yushi/BASF
alkyl poly(alkylene glycol) ether blend Bioquest 1110/1120a Baker Hughes
oxyalkylated alkylphenolic resins Bioquest 1395a Baker Hughes
organic oil Hodag K-4 UKb
UK Troy 333 UK
SAG 4130 UK
Clerol FBA 622 Cognis/Henkel Nopco
DL-2000 UK
a Antifoams offered commercially but not cited in fermentation literature at this time. b UK ) unknown.

cells, and not to interfere with electrodes such as pH or DO some cases, a balance is explored between selection of a costlier
sensors (Vardar-Sukan, 1992). but less utilized (and usually more effective) antifoam and
Speed/LongeWity. Antifoam effectiveness is measured by the selection of a cheaper but more highly utilized (and usually less
rate at which foam collapses and the duration of its action both effective) antifoam (Corbett, 1985).
initially upon addition and over the period it is present in the Solubility. Insolubility of antifoams is preferred to promote
broth (Byrant, 1970). Since often antifoams are added on their activity at low concentrations (Hall, 1971). Together with
demand, initial instantaneous action is highly desirable (Vardar-
low solubility, antifoams also should have a low critical micelle
Sukan, 1992) to achieve fast foam-breaking (Stanbury and
concentration to reduce foam stability through surface interac-
Whitaker, 1984, from Hall et al., 1973; Vardar-Sukan, 1992)
tions (Lee and Tynan, 1988). They should be insoluble in the
or knock down (Berovič, 1992; Solomons, 1967). Considering
its projected elaspsed time in the broth, a short-acting defoamer foaming medium, specifically possessing some hydrophobic
could be tolerated for fermentations that foam for only short character for low solubility but also having some hydrophilic
periods (Corbett, 1985). In general though, antifoams should character to ensure a low interfacial tension and thus a positive
be long-lasting and non-metabolizable (Stanbury and Whitaker, spreading coefficient (Vardar-Sukan, 1992; Sie and Schügerl,
1984, from Hall et al., 1973; Vardar-Sukan, 1992; Berovič, 1983). Determining where the antifoam resides in the fermenter
1992), as well as readily transferable to and dispersible in during a post-batch examination generates insight as to how it
fermentation broth (Stanbury and Whitaker, 1984, from Hall et works (Phillips et al., 1960) and its solubility. Specifically, for
al., 1973). spreading coefficients >0, air-oil-water systems are more
Concentration/Cost. Antifoams should be low cost and stable than air-water systems and oil likely exists as a thin
effective at low concentrations (Vardar-Sukan, 1992; Berovič, film at the air-water interface; when the spreading coefficient
1992; Stanbury and Whitaker, 1984, from Hall et al., 1973). In is <0, oils exists as discrete droplets in the bulk liquid (Weng
Biotechnol. Prog., 2007, Vol. 23, No. 4 775

Table 4. Selected Antifoams, Based on Fatty Acids/Esters (FE), Polyesters (PEP), and Oils (NE) Noted in Fermentation Literature
chemical tradename vendor
alkylpolyalcoxyester Clerol FBA 3107a Cognis
polyester polyol Desmophen 3600, 3900 Bayer AG
ester and polyalcohol Struktol SB2023 Schill and Seilacher
fatty acid ester O-30 Sigma
fatty acids, paraffin oil and non-ionic emulsifier Contraspum 210 Zschimmer and Schwarz
alkoxylated fatty acid esters on vegetable base Struktol J673 Schill and Seilacher
polyalkoxyester (EO/PO block polymer/ester) Biospumex 153 K Cognis
poly(ethylene glycol)-fatty acid ester and silicone oil AF emulsion Nakarai
glycerol alkyl oleate Atlas G 5600 Atlas Europol-ICI
natural oils various (lard, lard burning, soybean, corn, various
maize, cod liver, cottonseed (Proflo),
olive, sunflower, safflower, peanut,
ground nut, grape seed, linseed,
poppyseed, caster, palm
additives in natural oils 8% Alkaterge in grapeseed oil various
3% Alkaterge in lard oil
1% PPG in lard oil
octadecanol (stearyl alcohol)
1% octadecanol in ethyl alcohol
3% octadecanol in soybean oil, lard oil,
or peanut oil
substituted oxazolidine Alkaterge C commercial solvents
66 Alkaterge commercial solvents
cetyl polyoxyethylene condensate Lubrol Wa ICI
likely PAG-based, some containing fatty acid esters Glanapon Bussetti/Jastbolaget
non-silicone Nopco (Foam Master) Cognis/Henkel
UKb Prochem #51 Lambent
Dehysan 2111 Henkel
a Antifoams offered commercially but not cited in fermentation literature at this time. b UK ) unknown.

et al., 1997). In addition, non-metabolizable synthetic antifoams ingly, oil-based antifoams (e.g., polyalkyleneglycols, PAG)
are likely to coat vessel surfaces, thus requiring use of larger require longer sterilization times than silicon-in-water emulsion
amounts. (SE) antifoams (Solomons, 1969). The early literature describes
Chemical Properties and Toxicity. Antifoams should have a variety of sterilization conditions for antifoams: (1) steriliza-
low surface and interfacial tension (Vardar-Sukan, 1992) and tion at 160 °C for 2-3 h under dry heat conditions versus 127
should be non-volatile (Vardar-Sukan, 1992; Berovič, 1992). °C for 90 min versus direct steam for 2-3 h at 127 °C, all
Antifoams should be non-toxic to humans (i.e., without handling successfully applied (Bungay et al., 1960), (2) autoclaving for
hazards) and the environment (i.e., able to be sewered in 90 min at 125 or 127 °C (both temperatures cited, 127 °C
reasonable quantities), and especially non-toxic to the organism corresponded to 1.5 atm pressure) sufficient for 2 L quantities
being cultured (Stanbury and Whitaker, 1984, from Hall et al., but dry sterilization at 160 °C for 3 h inadequate (Paladino et
1973; Vardar-Sukan, 1992; Berovič, 1992). They should be non- al., 1954), (3) sterilization of larger than laboratory amounts
explosive and non-corrosive to vessels (either during sterilization routinely conducted by live steam injection at 2 atm pressure
or fermentation); they should have no or low flammability (Paladino et al., 1954). Small amounts of condensate formed
(Vardar-Sukan, 1992). during autoclaving or provided by live steam are deemed
Mass Transfer. There should be no appreciable adverse effect sufficient to provide the moist heat environment necessary for
on oxygen transfer by antifoams (Vardar-Sukan, 1992; Stanbury microorganism destruction at lower steam sterilization temper-
and Whitaker, 1984, from Hall et al., 1973). However, carbon atures (Paladino et al., 1954).
dioxide transfer as well as oxygen transfer is affected by the Sterilization of antifoams to be added to fermentations mid-
presence of antifoams with changes in the carbon dioxide cycle can be difficult, especially when they are undiluted, owing
removal rate causing changes in pH (Koch et al., 1995). to low vapor pressures and despite sometimes lower spore
Downstream Isolation. There should be little or no interfer- D-values in antifoam relative to water (Junker et al., 1999).
ence with subsequent product isolation or purification (So- [Specifically, the D-values of Antifoam C, UCON LB625, and
lomons, 1967; Stanbury and Whitaker, 1984, from Hall et al., P2000 relative to water are 0.8, 0.154, and 0.865, respectively
1973). They should be non-reactive with the product and not (Junker, 2001).] Interestingly, some antifoams are designed to
impart any odor or color to the isolated product (Vardar-Sukan, be added post-sterilization (i.e., they are not stable at sterilization
1992). temperatures) which means a second antifoam can be required
Assay. Selection of a specific antifoam also can be strongly during batching and sterilization. In other cases, antifoams (e.g.,
influenced by whether a straightforward and reliable assay is UCON HB) can be designed to become less water soluble as
available for its detection downstream or if a suitable assay can temperature increases (reverse solubility), thus becoming more
be developed readily. Two example methods are published: a effective antifoams in applications such as boiler feedwater
reverse phase LC method with evaporative light scattering (Currie, 1953). However, these novel antifoams are only
developed for simethicone (Moore et al., 2002) and a super- transferable to fermentation applications if they are biocom-
critical fluid chromatography method developed for UCON- patible.
LB625 (Dehaven et al., 1997). General Trends in Application. The selection of the
Sterilization. Antifoams should be stable during sterilization appropriate antifoam agent for a specific fermentation applica-
and able to be reliably steam-sterilized (Vardar-Sukan, 1992; tion must be “treated with caution and understanding” (Duitschae-
Stanbury and Whitaker, 1984, from Hall et al., 1973). Interest- ver, 1988). No single antifoam has all the properties for all foam
776 Biotechnol. Prog., 2007, Vol. 23, No. 4

control situations, and antifoam selection mostly is accomplished traceability. Often a single trade name may encompass several
by trial and error (Solomons, 1969). Often the only available types of antifoam of differing chemical composition so that the
method of selection is experimentation (Berovič, 1992), making product number becomes the critical differentiating factor. The
empirical tests of antifoams necessary for each individual system experimental sections of literature reports often have insufficient
(Schügerl, 1985). Interestingly, antifoams capable of destroying information regarding the antifoam employed (e.g., concentra-
foam in one application could stabilize foam in another tion, vendor, trade name and product number, chemical com-
application (Vardar-Sukan, 1991). position), and in some cases, antifoam is not used at small scale
Only a few authors indicate trends in antifoam usage (shake flasks or stirred tank bioreactors less than 20 L), or if it
depending on the fermentation process. Oily liquids such as is used, its description (e.g., type, amount, and addition
polyglycols are most effective in fungal fermentations, and conditions) is not specified. In other cases, antifoam is not
silicone-based antifoams are most effective in bacterial fermen- explicitly reported as a fermentation ingredient, although it
tations (Solomons, 1969). Specifically, PPG is effective at 0.2% appeared likely used from the experiment’s context. Since in
in a fermentation of Penicillium sp., but silicone emulsion is various organizations (e.g., academic, industrial, government),
not; PPG is not effective in a bacterial fermentation, but silicon typically one main antifoam is used for processes of the same
emulsion is effective (Solomons, 1967). Silicone defoamers nature, only one literature report from each group for the same
generally are not satisfactory in mold fermentations because they cultivation system was included in the analysis. Antifoams used
are inactivated by molds (Ghildyal et al., 1988; Solomons, 1967). with carriers (Berovič, 1992) or antifoams used synergistically
poly(dimethylsiloxane)s are most suitable for bacterial fermenta- for a greater combined effect (either mixed before addition or
tions at alkaline pH and for yeast fermentations regardless of added alternately) than when used individually (Berovič, 1992)
pH (Ghildyal et al., 1988). Often only 10 ppm or less of a poly- also are noted.
(dimethylsiloxane) is needed to be effective compared with 100 A sufficient number of published reports are examined to
to 1000 ppm for lower cost organic antifoams (Ross, 1967). determine if any trends in class of antifoam usage exist as
Soybean and lard oil antifoams disturb sugar utilization for summarized in Table 7. Regardless of culture type, for complex
sodium gluconate fermentation by A. niger, but octanol in ethyl media, a majority of antifoams utilized are PAG, with silicone/
alcohol does not (Blom et al., 1952). Finally, inert non- silicone emulsions forming a second significant grouping. For
metabolizable antifoams are preferred for enzyme production, semi-defined media, both PAG and silicone/silicone emulsions
but metabolizable antifoams are favored for secondary metabo- are most commonly selected, although fewer studies are
lite production (Schügerl, 1985). conducted using this media type. For defined media, primarily
There are some but comparatively few reports of antifoam PAG antifoams are utilized with far fewer selections of silicone/
use in polysaccharide fermentations or plant cell cultivations, silicone emulsions. By far the most reports were collected for
presumably owing to the high viscosity of these broths, which lab-scale bioreactors, although more studies are reported for pilot
reduces foaming. Antifoams used in plant cell culture are and production cultivations using complex media than other
sometimes used to control formation of crusts that prevent broth media types.
circulation (Bond et al., 1987). Silicon antifoam toxicity varies
In addition, a sufficient number of reports are examined where
according to the type of plant culture (Bond et al., 1987).
possible to acquire enough data to determine if any trends in
Similarly, only relatively few authors directly compare the amount of initial antifoam usage exist as summarized in
antifoams for their effect on production and other fermentation Table 8 using the same set of reports used for Table 7. Initial
parameters (Table 5). These studies are summarized with the antifoam amounts are greater for fungal and filamentous cultures
antifoam class listed according to Tables 2, 3, and 4 to facilitate grown on complex media compared with semi-defined and
study comparisons. In some cases, antifoams of the same class
defined media. For yeast, single cell bacterial and animal/other
have substantially different effects, suggesting that the propri-
cells, no such difference based on medium type is apparent.
etary nature of antifoam composition can be an important
Antifoam often is added initially to prevent foaming during
influence. In other cases, the antifoam class uniquely impacts
media preparation, transfer, and sterilization, as well as during
performance. When such studies are undertaken, some utilize
cultivation. Since foaming would be expected to decrease at
fermenters, whereas several others utilize shake flasks. Generally
laboratory scales, initial amounts likely might be skewed by
small amounts of most antifoam agents increase yield over no
the large fraction of laboratory reports in the set (Table 7).
antifoam addition in shake flask experiments since surface
oxygen transfer is presumed higher when no foam was present Types of Antifoams. Silicone-Based. Silicones are poly-
(Stefaniak et al., 1946). It can be difficult to apply shake flask (dialkylsiloxane)s, most commonly poly(dimethylsiloxane)s
results to predict bioreactor performance, however, since no (PDMS). Their typical viscosity is 60,000 cSt, and their typical
sparging is present in shake flasks. One interesting study average molecular weight is 270,000 (Moore et al., 2002; Hall,
attempts unsuccessfully to relate the fatty acid composition of 1971). They are virtually insoluble in water, possessing low
natural oils to its foam suppression capabililty (Table 6). volatility (Hall, 1971). Their surface tension is about 21 dyn/
Consequently, there does not appear to be known methods to cm, and their interfacial tension is about 42 dyn/cm (Hall, 1971).
predict an antifoam’s performance on the basis of its chemical These properties translate into low spreading coefficients and
composition. low dispersion in foaming systems, making them less effective
A tabulation of the antifoam class and specific antifoam type when used in their pure form (Hall, 1971).
is attempted for various fermentation process types to determine Consequently, silicone-based antifoams contain finely divided
if the fermentation literature suggests any further trends in solids, specifically silicas, creating high surface areas and an
antifoam selection (tabulation details not included). The most emulsifying agent to aid component distribution (Flannigan,
recently published similar tabulation of which this author is 1984). Some authors feel that hydrophobic, highly dispersed,
aware is nearly 20 years old (Ghildyal et al., 1988). Several solid silica particles are the main active ingredient and must be
factors make this analysis challenging. Many antifoam product present for silicone antifoams to break foam (Ghildyal et al.,
lines have changed vendors more than once, adversely impacting 1988). Simethicones are a complex mixture of high molecular
Biotechnol. Prog., 2007, Vol. 23, No. 4 777

Table 5. Effect of Antifoam Selection and Concentration on Production and Other Fermentation Parametersa
antifoam
process class specific antifoam parameter (units) amount scale reference
bioconversion PAG P2000 0 (titer, % control) 1 g/L SF Chartrain et al., 1990
of avermectin to SE FD62 91
27-OH avermectin PAG Pluronic L 122 23
by Nocardia PAG Pluracol P2010 18
polyoxamine Tetronic 901 21
NO lard burning oil 61 5 mL/L SF
Proflo oil 21
NO FD62 330 (titer, U/L) 15 g/L 14 L BR
NO lard burning oil 410 5 mL/L
lipase production N/A control 100 (titer, % control) 0 mL/L SF Marcin et al., 1993
by Pseudomonas SE FD 62 100 5 mL/L
aeruginosa S SAG 471 127
UK SAG 4130 94
PAG P2000 102
PAG SAG 5693 100
SE Mazu 210 S 94
SE Chemax DF-10 92
SE Chemax DF-30 102
NO cod liver oil 89 1 mL/L
corn oil 22
olive oil 89
soybean oil 80
lard oil 58
citric acid N/A control 56.8 (titer, g/L) no concn 1 L BR Kamal et al., 1999
production by S silicone oil 57.5 given
Aspergillus NO groundnut oil 73.8
ALC octadecanol 55.9
sodium gluconate NO soybean oil 0.12 (% glucose utiliization/h) 0.4 mL/L 1100 L BR Blom et al., 1952
by Aspergillus NO lard oil 0.04 0.15 mL/L
ALC 1% octanol in 0.86 15 mL/L
ethyl alcohol
anti-parasitic N/A control 17 (titer, mg/L) 0 mL/L SF Burg et al., 1979
production byStreptomyces PAG P2000 53 2.5 mL/L
anti-infective N/A none 90 (peak titer, mg/L) 0 SF Stefaniak et al., 1946
production by NO lard oil 112 1 mL/L
Penicillium NO soybean oil 98
ALC 3% octadecanol 95
in soybean oil
N/A none 97 0
ALC 3% octadecanol 101 1 mL/L
in lard oil
UK Nopco (type N/A) 121
NO corn oil 142
antifungal PAG P2000 -64 (titer, ( % max) 1 mL/L SF Sigmund and Hirsch, 1998
production by SE FD-62 +<10
Micromonospora S SAG-471 -<10
PAG UCON-LB625 +40
SE MAZU-DF201 -15
SE A/Sigma -<10
SE B/Sigma +30
SE C/Sigma +13
PAG-based Antifoam 204 +<10
dispersion
PAG UCON-LB625 47.6 (titer, mg/L) 0 mL/L SF
42.7 1 mL/L
69.1 3 mL/L
93.8 4 mL/L
PAG P2000 0 (titer, ( % max) 0 mL/L SF
-49 0.5 mL/L
-87 1.5 mL/L
-95 2.0 mL/L
baker’s yeast SE not given 100 (cell mass, g/L) 0.1 g/L not given Yasukochi, 1993
(cell mass) PAG 119
PAG 121
NO 108
recombinant PAG PPG 2000 0.004 (titer, U/L) 1 g/L 2 L BR Koch et al., 1995
protein production S S184 0.0019 2 g/L
by E. coli PAG/S VP1133 0.00224 1 g/L
SE SE9 0.00048 5 g/L
biopesticide PAG PPG 10,374 (titer, IU/µL) AAN 10 L BR Vidyarthi et al., 2000
production by SE A 11,690
Bacillus NO canola oil 12,348
peanut oil 13,020
olive oil 11,946
soybean oil 11,886
778 Biotechnol. Prog., 2007, Vol. 23, No. 4

Table 5. (Continued)
antifoam
process class specific antifoam parameter (units) amount scale reference
antibody SE w/PAG C plus 0.2% w/v 37 (titer, mg/L) 0.01 g/L 2 L BR Zhang et al., 1992
production by Pluronic F-68 40 0.05 g/L
hybridoma in 45 0.1 g/L
serum- containing 48 0.2 g/L
medium
human cell N/A control 1.0 (normalized viable cell) 0 ppm SF Meis et al., 2004
line (E1A SE A 1.6 100
complementing) SE B 2.15 100
SE SE15 1.6 200
PAG P2000 1.6 1000
PAG LB-625 0.3 1000
FE O-30 0 400
plant cell culture N/A control 12 (titer, g/L) 0 SF Wongsamuth and Doran, 1994
of Atropa SE C 12 0.3 g/L
PAG PPG 1025 12
PAG Pluronic PE6100 7.5
plant cell culture N/A control 4.2 (g/L dcw) 0 ppm 8 L BR Bond et al., 1987
of Catharanthus SE 1510 8.7 0.1 g/L
N/A N/A 2.8 0 ppm
PAG P2000 10.6 (g/L dcw) 0.1 g/L
plant cell S DC-A 99.4 (relative growth) 0.1 g/L SF Wang and Staba, 1963
culture of Mentha SE DC-B 99.3 0.1 g/L
SE FD-62 99.5 0.1 g/L
SE GE-60 98.2 0.3 g/L
UK Troy-333 99.1 0.1 g/L
NO safflower 109.7 0.1 g/L
a 15,000 ppm of FD62 required upon scale up to 14 L to control foam (appearing upon bioconversion substrate addition) which lowered titer. Lard oil

(5 mL/L in medium charge and 6 mL/L added mid-cycle) used for 500 L fermentations with aeration rate of 50 Lpm (Chartrain et al., 1990). Abbreviations:
PAG, poly(alkalene glycol); S, silicone (100%); SE, silicone emulsion; NO, natural oil; ALC, alcohol; PEP, polyester polyol; FE, fatty acid or fatty acid
esters including synthetic oils; OX, oxazolidine; UK, unknown (not enough information available either from vendors or from literature); SF, shake flask;
BR, bioreactor.

Table 6. Major Fatty Acid Composition of Natural Oilsa


saturated fatty acids (%) unsaturated fatty acids (%)
natural oil palmitic stearic arachidic palmitoleic oleic linoleic linolenic
soybean 8.0 6.0 0.6 0.2 22.0 52.0 8.0
cottonseed 24.0 2.0 0.3 0.9 21.0 50.0 0.1
linseed 6.5 5.5 0 0.2 21.0 15.0 51.0
olive 9.5 2.0 0.3 1.6 81.0 7.0 1.0
sunflower 6.0 4.0 0.6 0.2 36.0 54.0 0
sesame 8.0 6.0 0.5 0.4 40.0 42.0 0.3
a Listed in decreasing order of effectiveness in suppressing foam in unsterilized medium (Vardar-Sukan, 1988).

Table 7. Summary of Antifoam Selection and Fermentation Scale for Various Culture/Media Combinations Based on Published Fermentation
Studiesa

total reports antifoam class (% reports) scale (% reports)


culture media (class/scale) S SE PAG PEP NO FE ALC OX UK flask lab pilot prod UK
fungal complex 53/59 11 9 52 6 6 4 6 6 0 7 56 29 5 3
semi-defined 7/10 0 4 4 2 0 0 0 0 0 0 70 30 0 0
defined 12/15 8 8 60 8 8 8 0 0 0 0 80 20 0 0
filamentous bacterial complex 38/42 18 13 45 5 10 0 0 0 9 7 36 48 9 0
semi-defined 5/5 20 0 40 0 40 0 0 0 0 20 40 20 20 0
defined 9/9 0 11 89 0 0 0 0 0 0 11 78 11 0 0
yeast complex 12/14 17 17 58 0 0 8 0 0 0 7 57 36 0 0
semi-defined 10/10 20 40 30 0 0 10 0 0 0 0 70 30 0 0
defined 24/28 8 8 67 0 0 13 0 0 4 0 89 11 0 0
single cell bacterial complex 40/47 13 25 55 5 0 0 0 0 2 2 79 15 2 2
semi-defined 19/19 16 21 42 5 0 0 0 0 16 0 100 0 0 0
defined 34/40 6 3 85 0 0 0 0 0 6 2 73 18 5 2
animal and other cells complex 3/3 0 67 33 0 0 0 0 0 0 0 100 0 0 0
semi-defined 1/1 0 100 0 0 0 0 0 0 0 0 0 0 0 100
defined 3/3 0 67 33 0 0 0 0 0 0 0 33 67 0 0
a Higher number of scale reports due to some reports containing studies at multiple scales. Reports that compare multiple antifoams in Table 5 are omitted

from this analysis.

weight PDMS oligomers such as dimethicon with particulate oil-in-water emulsions of 10-30 vol % (Solomons, 1967),
silicon dioxide added (Moore et al., 2002). formulated to overcome problems with diluting antifoams
Mixtures of mineral or vegetable oils directly with silicone suitably for addition to fermentations (Keill et al., 1976). The
fluids and finely divided solids prove difficult to maintain end result of this emulsification with surfactants and stabilizers
homogeneous conditions prior to use (Flannigan, 1984). Con- is to raise hydrophilicity and improve dispersibility (Hall, 1971).
sequently, finely divided silica particles are dispersed in PDMS Water-based silicone emulsions can be less effective in control-
Biotechnol. Prog., 2007, Vol. 23, No. 4 779

Table 8. Summary of Antifoam Amount for Various Culture/Media Combinations Based on Published Fermentation Studiesa
antifoam amount (% reports)
total e0.01% 0.01<Xe0.1% 0.1<Xe0.2% 0.2<Xe1.0% 1.0<Xe2.0% >2%
culture media reports UK AANb e0.1 mL/L 0.1<Xe1 mL/L 1<Xe2 mL/L 2<Xe10 mL/L 10<Xe20 mL/L >20 mL/L
fungal complex 54 2 13 9 31 24 29 2 0
semi-defined 9 11 11 33 33 12 0 0 0
defined 13 7 46 8 39 0 0 0 0
filamentous complex 36 0 22 14 42 8 14 0 0
bacterial
semi-defined 5 0 20 0 60 0 20 0 0
defined 9 0 22 11 45 11 11 0 0
yeast complex 11 0 27 28 45 0 0 0 0
semi-defined 10 0 20 30 50 0 0 0 0
defined 26 0 30 15 35 8 8 0 4
single cell complex 40 5 30 25 30 5 3 0 2
bacterial
semi-defined 19 10 37 10 38 0 5 0 0
defined 35 6 34 17 40 0 3 0 0
animal and complex 3 0 0 33 67 0 0 0 0
other cells
semi-defined 1 0 0 0 100 0 0 0 0
defined 2 0 0 100 0 0 0 0 0
a Reports that compare multiple antifoams in Table 5 are omitted from this analysis. b Added as needed.

ling foam than pure silicone, although the emulsions can be in Table 6. Lard oil, composed mainly of glycerides of oleic
readily added to fermentations (Keill et al., 1976). These and stearic acids, is one frequently used natural oil antifoam
emulsions require some form of mixing to avoid settling during (Rolinson and Lumb, 1953). Recently, vegetable oils often are
prolonged stagnant periods with settling occurring faster at selected for products with food applications since they can be
higher dilutions. In addition, silicone emulsions tend to demul- made kosher and are non-animal-sourced. Oils used as antifoams
sify and break up after prolonged sterilization (Ghildyal et al., tend to oxidize forming aldehydes, carboxylic acids, peroxides,
1988; Solomons, 1967). and other compounds with negative effects on fermentation
PDMS-organic copolymers have been created containing performance, necessitating co-addition of non-toxic antioxidants
PDMS (MW g 2000) and organic PAG (polyoxypropylene of (Rakyta et al., 1980) or short expiry dates.
MW g 800 and polyoxypropylene/polyoxyethylene of MW g Oils increase the bubble size of foams, making them less
1500) components (Keill et al., 1976). stable (Rols and Goma, 1991). Their low hydrophilicity limits
Polyethers. Polyethers typically are composed exclusively of their spreading/dispersion abilities and their high viscosity limits
ethylene or propylene oxides or random and block copolymers their antifoaming capabilities (Hall, 1971; Vardar-Sukan, 1992).
of ethylene and propylene oxides to achieve a greater range of The best oil source for antifoam applications is defined as the
surface activities (Hall, 1971). Polyalkylene glycols are poly- source with the lowest concentration resulting in reasonably fast
ethers of molecular weights typically in the range of 400-2500 foam collapse (Vardar-Sukan, 1991). However, foam suppres-
(although the molecular weight of Pluronic F68 is 8400, which sion (preventing/inhibiting foam) as well as foam collapse
may account for its more modest effectiveness as an antifoam), (eliminating/breaking foam) must be considered (Vardar-Sukan,
with sufficient molecular weight to decrease solubility and thus 1991). All natural oils suppress foam formation (Jones and
control foam effectively (Dow; Lee et al., 1993). These Porter, 1998) to some degree, depending partly on the difference
antifoams have low solubility and reach gas-liquid interfaces between their surface tension and that of the broth (Ross, 1967).
by diffusion (Lee et al., 1993). Specifically at 20 °C, poly- Natural oils are energy-containing, metabolizable nutrients
(propylene glycol) solubility in water is 15.5 g/L and polypro- (Jones and Porter, 1998). Their addition to cultivations can
pylene/polyethylene (copolymer) solubility in water is 3.4 g/L increase product yields (Vardar-Sukan, 1992) or adversely affect
(Pollard et al., 2006). Lower molecular weight polymers are cultures (Ishida and Isono, 1952), particularly for genuses (such
too water-soluble for foam control (Hall, 1971). Viscosities of as Mucor, Aspergillus, and Penicillium) with high lipase
polyethers with molecular weights of 2000 are about 340 mPas activities (Ishida and Isono, 1952). For example, soybean oil is
(Bayer AG). more efficient as a foam breaker (short-lived application) than
One interesting report describes use of a mixture of PPGs of a foam inhibitor (long-lived application) due to its consumption
various molecular weights. Specifically, PPG1025 (BDH), PPG by the mold under cultivation (Weng et al., 1997).
2025 (BDH), and FoamMaster PPG (Henkel) are used in a ratio The effectiveness of natural oils in suppressing foam varies
of 2:2:1, diluted in water about 25% v/v, and then added to a with medium type and state (Vardar-Sukan, 1988). For 0.5-
fungal Fusarium fermentation in defined medium to final 1.0% v/v natural oil added to unsterilized soybean flour-
concentrations of 0.4-0.7% v/v (Wiebe et al., 2001). Another containing medium, foam height decreases in the following
report describes dispersion of larger amounts of PPG using a order: linseed > cotton seed > sesame > olive > soybean >
non-ionic surfactant, an “all-in-one” formulation, which “solu- sunflower (Vardar-Sukan, 1992). For various natural oils added
bilizes” the PPG (e.g., EEA-142 Henkel-Nopco) (Lee et al., to unsterilized soybean flour-containing medium, efficiency
1993). coefficients (based on oil cost and the amount required to reduce
Natural Oils. Natural oils are the earliest antifoams (Hall, foam to a minimum height) decrease in the following order:
1971), serving as one type of water-insoluble organic liquid soybean >> cotton seed >> linseed >> olive > sunflower >
(Vardar-Sukan, 1992). These oils generally are esters of glycerol sesame (Vardar-Sukan, 1992). For various natural oils added
with long chain saturated or unsaturated monobasic acids to sterilized soybean flour-containing medium, efficiency coef-
(triglycerides) along with some free fatty acids (Hall, 1971; ficients decrease in the following order: cotton seed >>> olive
Gutcho, 1973). Typical compositions of natural oils are shown >> soybean > sunflower > linseed > sesame (Vardar-Sukan,
780 Biotechnol. Prog., 2007, Vol. 23, No. 4

1992). Finally, when 0.1 or 0.2% v/v of any one of several constituents when possible (e.g., oat flour, tomato paste, animal
natural oils (e.g., caster, corn, linseed, poppy seed, soybean) extracts, meals) is the most effective way to reduce foaming
are added to either unsterilized or sterilized complex media without adding antifoam.
(either 5% sugar beet cosette or soybean flour), the foam Autoclaving, batch, or continuous sterilization potentially
collapse time decreases from minutes to seconds (Vardar-Sukan, causes foaming of fermentation media (Bailey and Ollis, 1977).
1991). During sterilization, an increase in sterilization hold temperature
Depending on fermenter agitation and aeration rates, natural from 110 to 130 °C (30 min sterilization time) of molasses-
oils increase, not decrease, volumetric mass transfer coefficients containing media causes the foaming coefficient (foam height
up to 90% when added at 0.25% v/v to S. cereVisiae in complex scaled by the ratio of foam disintegration to formation time) to
medium, recovering some of the rate reduction observed for double (Viesturs, 1982). Foaming of potato protein liquor at
amounts <0.25% v/v relative to when no oil was present (Liu pH 5.2 (Emsland-starke, GmbH, FRG; protein content 30%, ash
et al., 1994). Below that threshold, volumetric mass transfer 17%, potassium content 7%, reducing sugar 8%) increases 2000-
coefficients are found to decrease by up to 60%, with the biggest fold during sterilization at 120 °C for 30 min (Schügerl, 1985;
jump between 0 and 0.005% v/v (Liu et al., 1994). In this yeast Kotsaridu et al., 1983b). For a complex medium autoclaved
fermentation, 1% v/v olive oil controls foam as well as 0.1% without sugars (20 g/L Pharmamedia, 2.5 g/L corn starch, 10
v/v PPG, with the olive oil yielding a similar cell density but a g/L ammonium sulfate, 1.3 g/L potassium monophosphate, 15
much higher KLa (Liu et al., 1994). In Hansenula cultivation g/L salts), foaminess is low during the subsequent fermentation;
using defined medium, soy (soybean) oil is less effective as an when the same medium is autoclaved with 120 g/L lactose and
antifoam than Desmophen 3600; the foam does not disappear, 2.5 g/L glucose, foaminess is considerable (Koch et al., 1995).
but rather it exhibits increased bubble sizes and becomes less Foaminess during sterilization can be a function of steriliza-
stable (Voigt and Schügerl, 1981). tion temperature, hold time duration, pre-sterilization pH, and
Polyalcohols. Aliphatic alcohols, such as n-octanol, act as antifoam concentration. Foaminess decreases from 2000 to 60
antifoams and are attractive to use in fundamental studies since as hold temperature decreases from 120 °C to 80, and hold
they are comprised of a single molecular species rather than a duration decreases from 30 to 12.5 min (Kotsaridu et al., 1983b).
mixture; however, they are more volatile than higher molecular Foaminess decreases from 737 to 66 to 45 as pre-sterilization
weight polymers (Lee et al. 1993). Higher molecular weight pH decreased from 5.2 to 4.0 to 3.0, after sterilization at 120
alcohols such as octadecyl alcohol (also known as 1-octadecanol °C for 30 min (Kotsaridu et al., 1983b). Finally, foaminess
or stearyl alcohol) also possess antifoam properties but require decreases from 2000 to 67 as antifoam (Desmophen 3600)
a carrier since they are waxy solids. Early on, soybean oil, lard concentration increases from 0 to 5% after sterilization at 120
oil, or lard oil containing 1% octadecyl alcohol were successfully °C for 30 min (Kotsaridu et al., 1983b).
used with complex medium containing corn steep solids for Qualitative Pilot-Scale Foam Observations. Several seed
bacterial fermentations (Dworschack et al., 1954) and production media containing complex and in some cases
particulate media components were examined during and after
Specific Foaming Situations batch sterilization in pilot plant fermenters to qualitatively assess
Literature Reports. A few specific reports describing foaming characteristics.
foaming during cultivations as a function of media composition Batch sterilization is conducted by heating up medium using
are available. Selection of medium components and their jacket steam until a temperature of 95 °C is reached. At 95 °C,
concentrations notably affects foaming. In a very early study the jacket steam is discontinued and sparger steam is applied.
at the 30 L bioreactor scale with several fermentation systems, At 105 °C, steam is applied through the other internals (such
4-10-fold more antifoam was required for complex than for as the inoculation and medium transfer lines) until the temper-
synthetic media, but this comparison appears clouded since ature is a few degrees shy of the target hold temperature,
different cultures were used (i.e., no report of the same culture typically 122-123 °C. During the 40 min (seed vessels) or 45
being employed for both synthetic and complex media) (Rivett min (production vessels) sterilization hold period, the sparger
et al., 1950). Foaming during cultivations of Actinomyces in steam remains flowing, but steam through the other internals is
complex medium is minimized by varying media composition, discontinued. Temperature is controlled through small, manual
specifically using the same components at different relative changes in fermenter back-pressure. At the conclusion of the
concentrations (Soifer et al., 1974). Overall, the effect of sterilization, vessel back-pressure is increased to 1.5 kg/cm2 to
individual medium components on foaming requires further control foaming, sparger air is turned on at a low flowrate,
study (Ghildyal et al., 1988). followed by shutting off the sparger steam and applying cooling
In a model aqueous solution, soybean meal both strongly water to the vessel jacket. Once the temperature falls to below
foams and serves as foam stabilizer depending on its concentra- 40 °C, operating conditions are slowly adjusted to pre-
tion with its maximum ability to stabilize foam at 5 wt % and inoculation set points.
its foaming being lower above and below 5% (Szarka and In general, at the 600 L pilot scale, complex media have a
Magyar, 1969). Glycerol also acts as a foam stabilizer (Szarka tendency to foam during heat up, prior to reaching the desired
and Magyar, 1969). Furthermore, glucose concentrations from sterilization hold temperature. In cases where foam is somewhat
1% to 22% stabilize foam in 1-3% solutions of soybean meal controllable [Medium 1 ) 30 g/L Stadex 60 (dextrin, A.E.
(Szarka and Magyar, 1969). Foaminess of cell-free medium is Stalely, Tate and Lyle, Decatur, IL), 20 g/L Pharmamedia
believed caused mainly by the presence of corn steep liquor (cotton seed flour, Traders Protein, Memphis, TN), 2.5 g/L
(CSL) and a linear relationship between CSL concentration and Amberex pH (autolyzed yeast extract, Sensient Flavors, Indi-
foaminess exists (Burschapers et al., 2002). anopolis, IN), 7.5 g/L sucrose, 5.5 g/L cerelose (glucose
Additional reports are available describing foaming of media monohydrate), 2.0 mL/L P2000 (Dow, Freeport, TX)], a few
during sterilization or immediately afterward. Certain media inches of foam typically appears during heat up and/or when
ingredients cause foam upon sterilization, particularly complex sparger steam is applied. When internals are applied a few
or natural ingredients. The elimination of foam-forming medium minutes later, foam decreases and often dissipates only to
Biotechnol. Prog., 2007, Vol. 23, No. 4 781

Table 9. Qualitative Observations of Foaming during Various Stages of Batch Media Sterilization for Medium 3 in 600-L Pilot-Scale Vesselsa
medium heat up observations sterilization temp hold time observations cool down observations
all components when sparger steam applied, 122-124 °C foam gradually increased despite pressure increase
foam rose to top within a (more foam) to top, then subsided within to 1.5 kg/cm2 prior to
minute; foam decreased normal next 3-5 min;lowering introducing sparger air,
significantly shortly after pressure, even in small significant foaming
(<1 min) internals applied increments, results in observed
immediate foam
all components normal foam leaving 14 in.
foam to top when sparger foam to top during first
headspace
steam applied; minimal when 10 min, then subsided
all components 60 min foam to top, then subsided
applyng internals leaving 8′′ headspace
leaving 12 in. headspace
all components minimal foam observed normal minimal foam no foam
(including 2 mL/2000)
plus 20 mL/L soybean oil
a All sterilization hold times were 45 min and temperatures were 124-126 °C unless otherwise noted. Heat up and cool down procedures were normal.

reappear, sometimes rising considerably, when internals are shut Lpm. Interestingly, about 20-30 min after inoculation, the foam
off and to foam slightly during the sterilization hold phase. Foam drops dramatically (about 16 in.). As is characteristic of stable
then increases slightly when airflow is applied or in the best foams, this foam is primarily sensitive to surface-active
situations decreases or dissipates. components believed present in the inoculum.
In less optimal cases, foaming is less controllable, rising to
the vessel sightglass (top) at various points during the steriliza- Summary
tion process. For a complex medium [Medium 2 ) 20 g/L
The desired overall characteristics of antifoams are sum-
Stadex 60, 5 g/L Pharmamedia, 3 g/L NZ-Amine A (Quest
marized and presented as a compiled, comprehensive list of
International, Hoffman Estates, IL), 2 g/L yeast extract 106
properties, from which the most critical attributes must be
(autolyzed, BioSpringer USA, Minneapolis, MN), 5.5 g/L
selected for the specific process of interest. Consequently, trends
cerelose, 2 mL/L P2000], increasing the pre-sterilization P2000
in antifoam usage remain as general guidelines, based on
concentration to 6 mL/L or adding 10 mL/L soybean oil prevents
experiences of specific practitioners for their scale and cultures
foam rising to the top when internals are applied. Neither of
of interest. Despite the plethora of available antifoams, certain
these solutions is amenable for downstream processing, however.
antifoam classes more or less consistently are used for specific
In contrast, little effect is observed when 10 mL/L silicone oil
types of cultivations. Recent data from screening of alternative
(Dow-Corning 200, 10 cSt) is added. For another complex
antifoams at the shake flask and laboratory bioreactors also is
medium [Medium 3 ) 37 g/L Stadex 60, 5 g/L soy flour
presented, illustrating the process-specific nature of antifoam
(Thumb Oilseed, Ubly, MI), 7.9 g/L Hy-case (casein hydroly-
selection based on the desired process quantities for optimiza-
sate, Sheffield, Norwich, NY), 8.5 g/L yeast extract 106, 3.2
tion. By far the majority of foaming and antifoam studies are
mL/L P2000], foam is greater during cooldown for 60 min vs
conducted at the laboratory scale with significant attention
45 min sterilization hold times, consistent with more extensive
seldom devoted to finding an adequate scale down model for
Maillard reactions. In contrast to higher foam expected from
foam generation and prevention exploration. Some studies
greater Maillard degradation at higher temperatures, foam
addressing comparisons of antifoams and selection of medium
actually is lower during the sterilization hold time when the
components to minimize foam are conducted in ways that mimic
sterilization temperature is 124-126 °C versus 122-124 °C,
large-scale performance. Until further studies are available, the
presumably due to the higher vessel back-pressure of 1.3-1.35
strongest approach remains to reformulate media to minimize
kgf/cm2 required to attain the higher temperature. Table 9
or omit foaming components for processes that cannot tolerate
summarizes foaming observations during batch sterilization for
high foaming, high contamination rates, and/or high additions
Medium 3.
of antifoams.
In another complex medium at the 600 L scale [Medium 4
) 20 g/L Stadex 60, 3 g/L NZ amine A, 2 g/L yeast extract, Notation
5.5 g/L cerelose, 2 mL/L P2000] foam is watery/soapy in nature,
with about 6 in. appearing when sparger air is applied, a surface area per unit volume, 1/cm
disappearing when internals are applied, and then re-appearing DO dissolved oxygen, % sat
when internals are shut-off. A few inches remain during the KLa volumetric mass transfer coefficient, h-1
sterilization, there is no increase when sparger air is applied to OTR oxygen transfer rate, mmol/L-h
end the sterilization, and foam decreases as the batch cools, OUR oxygen uptake rate, mmol/L-h
possibly because the watery/soapy foam responds well to the Abbreviations
increased back-pressure applied during cooldown. Increasing
the PAG antifoam (2, 4, 8, and 16 mL/L) does not appreciably AAN added as needed
decrease foam; in fact foaming when air is applied to the AF antifoam
medium containing 16 mL/L is 2- to 3-fold greater. AFA antifoam agents
In another complex medium at the 15,000 L scale [Medium ALC alcohol
5 ) 6.5 g/L soy peptone SL (Marcor, Carlstadt, NJ), 6.5 g/L BR bioreactor
yeast extract 106, 5 g/L Hy-case SF, 2 mL/L P2000], foaming CSL corn steep liquor
is not problematic during sterilization. However, a stable foam DF defoamer
appears when the medium is cooled below 40 °C, which does EO ethylene oxide
not decrease with higher back-pressure or agitation rate but rises FCO foam control agent
about 6 in. when the airflow rate is raised from 1000 to 2000 FE fatty acid or fatty acid ester
782 Biotechnol. Prog., 2007, Vol. 23, No. 4

NO natural oil Bull, D. N.; Kempe, L .L. Influence of surface active agents on oxygen
OX oxazolidine absorption to the free interface in a stirred fermentor. Biotechnol.
Bioeng. 1971, 13, 529-547.
PAG poly(alkylene glycol)
Bumbullis, W.; Kalischewski, K.; Schügerl, K. Foam behavior of
PDMS poly(dimethylsiloxane)s biological media, II. Salt effects. Eur. J. Appl. Microbiol. Biotechnol.
PEP polyester polyol 1979, 7, 147-154.
pI protein isoelectric point Bumbullis, W.; Schügerl, K. Foam behavior of biological media, V.
PO propylene oxide Alcohol effects. Eur. J. Appl. Microbiol. Biotechnol. 1979, 8, 17-
PPG poly(propylene glycol) 25.
Bungay, H. R.; Simons, C. F.; Hosler, P. Handling of antifoam oils for
S silicone (100%)
fermentations. J. Biochem. Microbiol. Technol. Eng. 1960, 2 (2),
SAFD stirring as foam disruption 143-155.
SE silicone emulsion Burg, R. W.; Miller, B. M.; Baker, E. E.; Birnbaum, J.; Currie, S. A.;
SF shake flask Hartman, R.; Kong, Y. Monaghan, R. L.; Olson, G.; Putter, I.; Tunac,
UK unknown J. B.; Wallick, H.; Stapley, E. O.; Oiwa, R.; Omura, S. Avermectins,
XFO X-foam new family of potent anthelmintic agents: Producing organism and
fermentation. Antimicrob. Agents Chemother. 1979, 15 (3), 361-
367.
Burschapers, J.; Schustolla, D.; Schugerl, K.; Roper, H.; de Troost-
References and Notes embergh, J. C. Engineering aspects of the production of sugar
Abdullah, M. A.; Ariff, A. B.; Marziah, M.; Ali, A. M.; Lajis, N. H. alcohols with the osmophilic yeast Moniliella tomentosa Var pollinis,
Strategies to overcome foaming and wall-growth during the cultiva- Part 2. Batch and fed-batch operation in bubble column and airlift
tion of Morinda elliptica cell suspension culture in a stirred-tank tower loop if reactors. Process Biochem. 2002, 38, 559-570.
bioreactor. Plant Cell, Tissue Organ Cult. 2000, 60, 205-212. Chain, E. B.; Gualandi, G.; Morisi, G. Aeration studies. IV. Aeration
Adler, I.; Diekmann, J.; Hartke, W.; Hecht, V.; Rohn, F.; Schügerl, K. conditions in 3000-liter submerged fermentations with various
Bubble coalescence behaviour in biological media II. Effect of microorganisms. Biotechnol. Bioeng. 1966, 8, 595-619.
antifoam additives. Eur. J. Appl. Microbiol. Biotechnol. 1980, 10, Chartrain, M.; White, R.; Goegelman, R.; Gbewonyo, K.; Greasham,
171-186. R. Bioconversion of avermectin into 27-OH avermectin. J. Ind.
Adler, I.; Fiechter, A. Valuation of bioreactors for low viscous media Microbiol. 1990, 6, 279-284.
and high oxygen transfer demand. Bioprocess Eng. 1986, 1 (2), 51- Chen, J.; Wang, H. Y. Bioprocess monitoring of dissolved oxygen using
59. a computerized pulsing membrane electrode. Biotechnol. Prog. 1993,
Ash, M.; Ash, I. Handbook of Industrial Surfactants, 3rd ed.; Vol. 1. 9, 75-80.
Synapse Information Resources, Inc.; Endicott: New York, 2000a; Chisti, Y. Animal cell culture in stirred bioreactors: Observations on
Vol. 1. scale-up. Process Biochem. 1993, 28, 511-517.
Ash, M.; Ash, I. Handbook of Industrial Surfactants, 3rd ed.; Synapse Christiano, S. P.; Fey, K. C. Silicone antifoam performance enhance-
Information Resources, Inc.; Endicott: New York, 2000b; Vol. 2. ment by nonionic surfactants in potato medium. J. Ind. Microbiol.
Atkinson, B.; Mavituna, F. Biochemical Engineering and Biotechnology Biotechnol. 2003, 30, 13-21.
Handbook; Nature Press: New York, 1983; pp 768-771. Corbett, K. Design, preparation and sterilization of fermentation media.
Bailey, J.; Ollis, D. F. Biochemical Engineering Fundamentals; In Moo-Young, M., Ed.; ComprehensiVe Biotechnology; Pergamon
McGraw-Hill, Inc.: New York, 1977; pp 553-554. Press: New York, 1985; pp 127-139.
Bartholomew, W. H.; Koslow, D. Automatic antifoam and nutrient feed Currie, C. C. Chemical antifoaming agents. In Bikerman, J. J., Perri,
control for bench scale fermentation. Ind. Eng. Chem. 1957, 49 (8), J. M., Booth, R. B., Currie, C. C., Eds.; Foams: Theory and
1221-1222. Industrial Applications; Reinhold Publishing Corporation: New
Berovič, M. Foam problems in fermentation processes. In Mukherjee, York, 1953; pp 297-329.
R. N. Downstream Process Biotechnology, Proc. Int. Semin. Meeting; de Haut, C. 2001. Foam control solutions in fermentation. Cognis
Tata-McGraw-Hill: New Delhi, 1992; pp 248-261. literature 2001;1-13.
Berovič, M.; Cimerman, A. Foam in submerged citric acid fermentation Dehaven, P. B.; Rowe, G.; Stambaugh, T.; Donnelly, J.; Hennessey, J.
on beet molasses. Eur. J. Appl. Microbiol. Biotechnol. 1979, 7, 313- The detection of UCON-LB625 in manufacturing process streams
319. using supercritical fluid chromatography-antifoam detection during
Blom, R. H.; Pfeifer, V. F.; Moyer, A. J.; Traufler, D. H.; Conway, H. process validation for vaccine adjuvant production by Neisseria
F.; Crocker, C. K.; Farison, R. E.; Hannibal, D. V. Sodium gluconate meningitis. Abstracts of Papers of the 213th ACS Meeting, Pt. 1;
production. Ind. Eng. Chem. 1952, 44 (2), 435-440. American Chemical Society: Washington, DC, 1997.
Bond, P. A.; Hegarty, P.; Scragg, A. H. 1987. The use of antifoams in Deindoerfer, F. H.; Gaden, E. L. Effects of liquid physical properties
the mass cultivation of plant cells. In Neijseel, O. M., van der Meer, on oxygen transfer in penicillin fermentation. Appl. Microbiol. 1955,
R. R., Luyben, K. Ch. A. M., Eds.; Proceedings of the 4th European 3, 253-257.
Congress on Biotechnology; Elsevier: Amsterdam, 1987; Vol. 2,
Denkov, N. D.; Cooper, P.; Martin, J.-Y. Mechanisms of action of
pp 440-443.
mixed solid-liquid antifoams. 1. Dynamics of foam film rupture.
Boon, L. A.; Hoeks, F. W. J. M. M.; van der Lans, R. G. J. M.; Bujalski,
Langmuir 1999, 15, 8514-8529.
W.; Nienow, A. W. Instabilities when using a standard (T/3) Rushton
turbine for stirring as foam disruption (SAFD). Can. J. Chem. Eng. Duitschaever, C.L.; Buteau, C.; Kamel, B.S. An investigation on the
2000, 78, 884-891. efficiency of antifoaming agents in aerobic fermentation. Process
Boon, L. A.; Hoeks, F. W. J. M. M.; van der Lans, R. G. J. M.; Bujalski, Biochem. 1988, 23 (6), 163-165.
W.; Wolff, M. O.; Nienow, A. W. Comparing a range of impellers Dworschack, R. G.; Lagoda, A. A.; Jackson, R. W. Fermentor for small-
for “stirring as foam disruption”. Biochem. Eng. J. 2002, 10, 183- scale submerged fermentations. Appl. Microbiol. 1954, 2, 190-197.
195. Edwards, M.; Eschenbruch, R.; Molan, P. C. Foaming in winemaking
Brown, A. K.; Isbell, C.; Gallagher, S.; Dodd, P. W.; Varley, J. I. A technique for the measurement of foaming in winemaking. Eur
Improved measurement and control of fermentation foams. Trans. J. Appl. Microbiol. Biotechnol. 1982, 16, 105-109.
Inst. Chem. Eng. 2001, 79 (C), 59. Elander, R. P. Bioprocess technology in industrial fungi. In Neway, J.
Brown, A. K.; Isbell, C.; Gallagher, S.; Dodd, P. W.; Varley, J. An O., Ed.; Fermentation Process DeVelopment of Industrial Organisms;
improved method for controlling foams produced within bioreactors. Marcel Dekker: New York, 1989; pp 186, 204, 209.
Trans. Inst. Chem. Eng. 2001, 79 (C), 114-120. Evans, J. I.; Hall, M. J. Foams and antifoams in fermentation. Process
Bryant, J. Anti-foams agents. In Norris, J. R., Ribbons, D. W., Eds.; Biochem. 1971, 63, 23-26.
Methods in Microbiology; Academic Press: New York, 1970; pp Flannigan, W. T. Compositions for the control of unwanted foam and
187-203 . their use. U.S. Patent 4,451,390, 1984.
Biotechnol. Prog., 2007, Vol. 23, No. 4 783

Flickinger, M. C.; Greenstein, M.; Bremmon, C.; Conlin, J. Strain Kotsaridu, M.; Gehle, R.; Schügerl, K. Foam behavior of biological
selection, medium development and scale-up of toyocamycin media, IX. pH and salt effects. Eur. J. Appl. Microbiol. Biotechnol.
production by Streptomyces chrestomyceticus. Bioprocess Eng. 1990, 1983a, 18, 60-63.
5, 143-153. Kotsaridu, M.; Müller, B.; Pfanz, V.; Schügerl, K. Foam behavior of
Gaden, E. L.; Kevorkian, V. Foams in chemical technology. Chem. biological media, X. Influence of the sterilization conditions on the
Eng. 1956, 63, 173-179. foaminess of PPL solutions. Eur. J. Appl. Microbiol. Biotechnol.
Garrett, P. R. 1993. The mode of action of antifoams. In Garrett, P. 1983b, 17, 258-260.
R., Ed.; Defoaming Theory and Industrial Applications; M. Dek- Kovalev, V. N.; Ivankova, T. A.; Bylinkina, E. S. Effect of chemical
ker: New York, 1993; Surfactant Science Series, Vol. 45, pp 1-117. foam suppressors on mass exchange rate in oxytetracycline biosyn-
Getchell, J. R. Instrumentation and control systems. In Vogel, H. C., thesis. Antibiotiki 1982, 27, 263-269.
Ed.; Fermentation and Biochemical Engineering Handbook: Prin- Lee, J. C.; Salih, M. A.; Sebai, N. N.; Withey, A. Control of foam in
ciples, Process Design and Equipment; Noyes Publications: New bioreactors-action of antifoams. In Nienow, A. W.; Ed.; Proceedings
Jersey, 1983; pp 400-429. of the 3rd International Conference on Bioreactors and Bioprocess
Ghildyal, N. P.; Lonsane, B. K.; Karanth, N. G. Foam control in Fluid Dynamics; MEP Ltd.: London, 1993; pp 275-287.
submerged fermentation: State of the art. AdV. Appl. Microbiol. Lee, J. C.; Tynan, K. J. Antifoams and their effects on coalescence
1988, 33, 173-222. between protein-stabilised bubbles. In King, R., Ed.; 2nd Interna-
Gutcho, S. J. Chemicals by Fermentation; Noyes Data Corporation: tional Conference on Bioreactor Fluid Dynamics; Elsevier Applied
New Jersey, 1973; pp 27-57, 287-288. Science Publishers: New York, 1988; pp 353-377.
Hall, M. J. Foams and foam control in fermentation processes. Prog.
Lengyel, Z. L.; Nyiri, L. Studies on automatically aerated biosynthetic
Ind. Microbiol. 1971, 12, 171-231.
processes. II. Occurrence and elimination of CO2 during penicillin
Hall, M. J.; Dickinson, S. D.; Pritchard, R.; Evans, J. I. Foams and
biosynthesis. Biotechnol. Bioeng. 1966, 8, 337-352.
foam control in fermentation processes. Prog. Ind. Microbiol. 1973,
Liew, M. K. H.; Fane, A. G.; Rogers, P. L. Fouling of microfiltration
12, 170-234.
membranes by broth-free antifoam agents. Biotechnol. Bioeng. 1997,
Hastings, J. J. H. Problems of biochemical engineering. Trans. Inst.
56 (1), 89-98.
Chem. Eng. 1954, 32, 11-22.
Hoeks, F. W. J. M. M.; Boon, L. A.; Studer, F.; Wolff, M. O.; van der Liu, H. S.; Chiung, W. C.; Wang, Y. C. Effect of lard oil, olive oil and
Schot, F.; Vrabel, P.; van der Lans, R.; Bujalski, W.; Manelius, A.; castor oil on oxygen transfer in an agitated fermentor. Biotechnol.
Blomsten, G.; Hjorth, S.; Prada, G.; Luyben, K.; Nienow, A. W. Tech. 1994, 8 (1), 17-20.
Scale-up of stirring as foam disruption (SAFD) to industrial scale. Meis, P.; Matallo, C.; Peltier, J.; Zhou, W.; Blumentals, M.; Amanullah,
J. Ind. Microbiol. Biotechnol. 2003, 30, 118-128. A. Compatibility studies for the use of chemical antifoams in the
Hoeks, F. W. J. M. M.; Van, Wees-Tangerman, C.; Gasser, K.; production of adenovirus suspension cultures. AIChE Annnual
Mommers, H. M.; Schmid, S.; Luyben, K. Stirring as foam disruption Meeting, Austen TX, 2004.
(SAFD) technique in fermentation processes. Can. J. Chem. Eng. Moore, D. E.; Liu, T. X.; Miao, W. G.; Edwards, A.; Elliss, R. A RP-
1997, 75, 1018-1029. LC method with evaporative light scattering detection for the assay
Ishida, Y.; Isono, M. Effect of antifoaming oils on penicillin fermenta- of simethicone in pharmaceutical formulations. J. Pharm. Biomed.
tion, III. Inhibition of oil-decomposed substances. J. Antibiot. 1952, Anal. 2002, 30, 273-278.
7, 381-387. Morão, A.; Maia, C. I.; Fonseca, M. M. R.; Vasconcelos, J. M. T.;
Ishizuka, H.; Wako, K.; Kasumi, T.; Sasaki, T. Breeding of a mutant Alves, S. S. Effect of antifoam additon on gas-liquid mass transfer
of Aureobasidium sp. with high erythritol production. J. Ferment. in stirred fermenters. Bioprocess Eng. 1999, 20 (2), 165-172.
Bioeng. 1989, 68 (5), 310-314. Motoki, M.; Okiyama, A.; Nonaka, M.; Tanaka, H.; Uchio, R.;
Jenkins, D.; Richard, M. G.; Daigger, G. T. Manual on the Causes Matsuura, A.; Ando, H.; Umeda, K. Transglutaminase. U.S. Patent
and Control of ActiVated Sludge Bulking and Foaming, 2nd ed.; 5,165,956, 1992.
Lewis Publishers: Ann Arbor, 1993; pp 8-11, 145-170. Noble, I.; Collins, M.; Porter, N.; Pyle, D. C.; Varley, J. Foaming in
Jewell, J. B.; Dunphy, G. B. Antifoaming agent produced by strains of fermentation: The biochemical basis. ICHEME-AdVances in Bio-
the entomopathogenic bacterium Xenorhabdus nematophilus and its chemical Engineering; Institute of Chemical Engineers: Rubgy,
effect on the development of the insect pathogenic nematode U.K., 1994a; pp 17-19.
Steinernema carpocapsae DD136. J. Gen. Appl. Microbiol. 1996, Noble, I.; Collins, M.; Porter, N.; Varley, J. An investigation of the
42, 39-49. physico-chemical basis of foaming in fungal fermentations. Bio-
Jha, B. K.; Christiano, S. P.; Shah, D. O. Silicone antifoam perfor- technol. Bioeng. 1994b, 44, 801-807.
mance: Correlation with spreading and surfactant monolayer pack- Olivieri, R.; Forconi, L.; Bianciardi, S.; Rappuoli, R. Foam formation
ing. Langmuir 2000, 16 (26), 9947-9954. and control in industrial fermentative production of modified
Jones, A. M.; Porter, A. M. Vegetable oils in fermentation: Beneficial diphtheria toxin by a mutant of Corynebacterium
effects of low level supplementation. J. Ind. Microbiol. Biotechnol. diphtheriae. Chim. Oggi 1993, 11 (1-2), 41-42.
1998, 21, 203-207. Paladino, S.; Ugolini, F.; Chain, E. B. Fermenters of 90 and 300 L
Junker, B. Technical evaluation of the potential for streamlining of capacity for vortex and sparger aeration. Rend. Ist. Super. Sanita
equipment validation for fermentation applications. Biotechnol. 1954, 17, 88-120.
Bioeng. 2001, 74 (1), 49-61.
Pandit, J. Foaming in biochemical reactors. ICHEME-AdVances in
Junker, B. H. Sterilization-in-place of concentrated nutrient solutions.
Biochemical Engineering; Institute of Chemical Engineers: Rugby,
Biotechnol. Bioeng. 1999, 62 (5), 501-508.
U.K., 1989; pp 45-62.
Kamal, K. P.; Verma, U. N.; Nag, A. K.; Singh, S. P. Effect of some
Paul, E. L.; Kaufman, A.; Sklarz, W. A. An industrial approach to
antifoam agents and oxygen transfer rate on citric acid production
integrated fermentation/isolation process development. Ann. N.Y.
by submerged fermentation. Asian J. Chem. 1999, 11 (3), 1020-
Acad. Sci. 1981, 369, 181-186.
1022.
Kawase, Y.; Moo-Young, M. The effect of antifoam agents on mass Pelton, R. A review of antifoam mechanisms in fermentation. J. Ind.
transfer in bioreactors. Bioprocess Eng. 1990, 5, 169-173. Microbiol. Biotechnol. 2002, 29, 149-154.
Keil, J. W. Foam control composition. U.S. Patent 3,984,347, 1976. Pfeifer, V. F.; Heger, E. N. Electronic foam controller for fermentors.
Koch, V.; Rüffer, H. M.; Schügerl, K.; Innertsberger, E.; Menzel, H.; Appl. Microbiol. 1957, 5, 44-47.
Weis, J. Effect of antifoam agents on the medium and microbial Pfeifer, V. F.; Vojnovich, C.; Heger, E. N. Itaconic acid by fermentation
cell properties and process performance in small and large reactors. with Aspergillus terreus. Ind. Eng. Chem. 1952, 44 (12), 2975-
Process Biochem. 1995, 30 (5), 435-446. 2980.
Koller, K. Foam control in fermentation processes. Chem. Eng. 2004, Pharmatec. Frustrating Foam. PharmaTEC International. Process World
111 (8), 24-27. Wide-PharmaTech, 2004; 2.
König, B.; Kalischewski, K.; Schügerl, K. Foam behavior of biological Phillips, K. L.; Spencer, J. F. T.; Sallans, H. R.; Roxburgh, J. M. Effect
media III. Penicillium chrysogenum cultivation foam. Eur. J. Appl. of antifoam agents on oxygen transfer in deep tank fermentations.
Microbiol. 1979, 7, 251-258. J. Biochem. Microbiol. Technol. Eng. 1960, 2 (1), 81-91.
784 Biotechnol. Prog., 2007, Vol. 23, No. 4

Pollard, J. M.; Shi, A. J.; Goklen, K. E. Solubility and partitioning Van’t Riet, K.; Tramper, J. Basic Bioreactor Design; Marcel Dekker,
behavior of surfactants and additives used in bioprocesses. J. Chem. Inc.: New York, 1991; pp 274-292.
Eng. Data 2006, 51 (1), 230-236. Van’t Riet, K.; Van Sonsbeek, H. M. Foaming, mass transfer, and
Prins, A.; van’t Riet, K. Proteins and surface effects in fermentation: mixing: Interrelations in large scale fermentors. In Ladisch, M. R.,
Foam, antifoam, and mass transfer. Trends Biotechnol. 1987, 5 (11), Bose, A., Eds.; Harnessing Biotechnology for the 21st Century;
296-301. American Chemical Society: Washington, DC,1992; pp 189-192.
Rakyta, Y.; Frimm, R.; Velvard, L.; Latsko, L.; Lukashikova, E. Vardar-Sukan, F. Effects of natural oils on foam collapse in biopro-
Antioxidant stabilization of antifoaming agents used in fermentation. cesses. Biotechnol. Lett. 1991, 13 (2), 107-112.
Antibiotiki 1980, 25, 12-16. Vardar-Sukan, F. Efficiency of natural oils as antifoaming agents in
Reisman, H. G., Ed.; Economic Analysis of Fermentation Processes; bioprocesses. J. Chem. Technol. Biotechnol. 1988, 43, 39-47.
CRC Press: Boca Raton, FL, 1988; pp 17, 26, 29, 34, 37-39, 52- Vardar-Sukan, F. Foaming and its control in bioprocesses. In Vardar-
54, 150-174. Sukan, F., Suha Sukan, S., Eds.; Recent AdVances in Biotechnology;
Rivett, R. W.; Johnson, M. J.; Peterson, W. H. Laboratory fermentor Kluwer Academic Publishers: Netherlands, 1982; pp 113-146.
for aerobic fermentations. Ind. Eng. Chem. 1950, 42 (1), 188-190. Viesturs, U. E.; Kristapsons, M. Z.; Levitans, E. S. Foam in micro-
Rolinson, G. N.; Lumb, M. The effect of aeration on the utilization of biological processes. In Gutschick, V. P., Harder, A., Humphrey,
respiratory substrates by Penicillium chrysogenum in submerged A., Kristapsons, M. Z., Levitans, E. S., Roels, J. A., Rolz, C.,
culture. J. Gen. Microbiol. 1953, 8, 265-272. Viesturs, U. E., Eds.; Microbes and Engineering Aspects. Springer-
Rols, J. L.; Goma, G. Enhanced oxygen transfer rates in fermentation Verlag: Berlin, 1982; pp 170-230.
using soybean oil-in-water dispersions. Biotechnol. Lett. 1991, 13 Vidyarthi, A. S.; Desrosiers, M.; Tyagi, R. D.; Valéro, J. R. Foam
(1), 7-12. control in biopesticide production from sewage sludge. J. Ind.
Romualdo, R.; Luiz, C. C.; Fernandes, L. A. Defoamer and methods Microbiol. Biotechnol. 2000, 25, 86-92.
of use thereof. U.S. Patent 6,448,298 B1, 2002. Vogel, H. C., Ed.; Fermentation and Biochemical Engineering Hand-
Ross, S. Mechanisms of foam stabilization and antifoaming action. book; Noyes: Park Ridge, NJ, 1983.
Chem. Eng. Prog. 1967, 63 (9), 41-47.
Voigt, J.; Schügerl, K. Comparison of single- and three-stage tower
Schubert, J.; Wan, L.; Lubbert, A. Foam suppression by bioreactor
loop reactors. Eur. J. Appl. Microbiol. Biotechnol. 1981, 11, 97-
retrofitting. In Nienow, A. W. Proceedings of 3rd International
105.
Conference on Bioreactor and Bioprocess; MEP Ltd.: London, 1993.
Vraná, D.; Seichert, L. Cytomophological comparison of mechanical
Schügerl, K. Foam formation, foam suppression, and the effect of foam
and chemical defoaming of a yeast culture. Folia Microbiol. 1988,
on growth. Process Biochem. 1985, 20 (4), 122-123.
33, 144-147.
Schügerl, K. On-line analysis and control of production of antibiotics.
Anal. Chim. Acta 1988, 213, 1-9. Wang, C.-J.; Staba, E. J. Peppermint and spearmint tissue culture II:
Sie, T.-L.; Schügerl, K. Foam behavior of biological media. XI. Dual-carboy cultures of spearmint tissues. J. Pharm. Sci. 1963, 52
Efficiency of antifoam agents with regard to their foam suppression (11), 1058-1062.
effect on BSA solutions. Eur. J. Appl. Microbiol. Biotechnol. 1983, Weng, C. J.; Ku, C. H.; Chen, Z. W.; Chen, T. L. Effect of soybean
17, 221-226. oil on antifoaming, oxygen transfer and penicillin productivity. J.
Sigmund, J. M.; Hirsch, C. F. Fermentation studies of rustmicin Chin. Inst. Chem. Eng. 1997, 28 (5), 307-311.
production by a Micromonospora sp. J. Antibiot. 1998, 51 (9), 829- Wiebe, M. G.; Robson, G. D.; Shuster, J.; Trinci, A. P. J. Evolution of
836. a recombinant (glucoamylase-producing) strain of Fusarium Venena-
Soifer, R. D.; Gorskaya, S. V.; Ivankova, T. A. Foaming as a control tum A3/5 in chemostat culture. Biotechnol. Bioeng. 2001, 73, 146-
factor of fermentation. Biotechnol. Bioeng. Symp. 1974, 4, 755- 156.
767. Wilde, P. J.; Husband, F. A.; Cooper, D.; Ridout, M. J. Destabilization
Solomons, G. L. Materials and Methods in Fermentation; Academic of beer foam by lipids: Structural and interfacial effects. J. Am.
Press: New York, 1969; pp 133-149. Soc. Brew. Chem. 2003, 61 (4), 196-202.
Solomons, G. L. Anti-foams. Process Biochem. 1967, 2, 47-48. Wolfes, H.; Schügerl, K. Foam behavior of biological media VIII.
Solomons, G. L. Aeration and agitation. Process Biochem. 1966, 1, Surface properties. Eur. J. Appl. Microbiol. Biotechnol. 1983, 17,
307-312,317. 371-375.
Stanbury, P. F.; Whitaker, A.; Hall, S. J. Principles of Fermentation Wongsamuth, R.; Doran, P. M. Foaming and cell flotation in suspended
Technology, 2nd ed.; Elsevier Science Ltd.: New York, 1995; pp plant cell cultures and the effect of chemical antifoams. Biotechnol.
211, 267. Bioeng. 1994, 44, 481-488.
Stanbury, P. F.; Whitaker, A. Principles of Fermentation Technology, Yagi, H.; Yoshida, F. Desorption of carbon dioxide from fermentation
1st ed.; Pergamon Press: Oxford, 1984. broth. Biotechnol. Bioeng. 1977, 19, 801-819.
Stefaniak, J. J.; Gailey, F. B.; Jarvis, F. G.; Johnson, M. J. The effect Yagi, H.; Yoshida, F. Oxygen absorption in fermenters-Effects of
of environmental conditions on penicillin fermentations with Peni- surfactants, antifoaming agents, and sterilized cells. J. Ferment.
cillium chrysogenum X-1612. J. Bacteriol. 1946, 52, 119-127. Technol. 1974, 52 (12), 905-916.
Su, W. W. Bioprocessing technology for plant cell suspension cultures. Yamashita, S. Instrumentation and control in recent fermentation
Appl. Biochem. Biotechnol. 1995, 50, 189-230. processes. pp. 441-463. In Terui, G., Ed.; Fermentation Technology
Szarka, L.; Magyar, K. The foams of fermentation broths, I. Some Today; Proc. IV IFS; Society of Fermentation Technology: Osaka,
parameters of the foaming of fermentation media. Biotechnol. Bioeng. 1972; pp 441-463.
1969, 11, 701-710. Yasukawa, M.; Onodero, M.; Yamagiwa, K.; Ohkawa, A. Gas holdup,
Takahashi, H.; Yoshida, F. Oxygen transfer in a bubble column for power consumption, and oxygen absorption coefficient in a stirred-
fermentation with Aspergillus niger. J. Ferment. Technol. 1979, 57 tank fermentor under foam control. Biotechnol. Bioeng. 1991, 38,
(4), 349-356. 629-636.
Takesono, S.; Onodera, M.; Ito, A.; Yoshida, M.; Yamagiwa, K.; Yasukochi, T. Antifoaming agent in fermentation industry. Yukagaku.
Ohkawa, A. Mechanical control of foaming in stirred-tank reactors. 1993, 42 (10), 792-799.
J. Chem. Technol. Biotechnol. 2001, 76 (4), 355-362. Zhang, S.; Handa-Corrigan, A.; Spier, R. E. Foaming and media
Taticek, R. A.; Moo-Young, M.; Legge, R. L. The scale-up of plant surfactant effects on the cultivation of animal cells in stirred and
cell cultures: Engineering considerations. Plant Cell, Tissue Organ sparged bioreactors. J. Biotechnol. 1992, 25, 289-306.
Cult. 1991, 24, 139-158.
van der Pol, L. A.; Bonarius, D.; van de Wouw, G.; Tramper, J. Effect Received January 30, 2007. Accepted May 14, 2007.
of silicone antifoam on shear sensitivity of hybridoma cells in sparged
cultures. Biotechnol. Prog. 1993, 9, 504-509. BP070032R

You might also like