Sie sind auf Seite 1von 8

APPLIED AND ENVIRONMENTAL MICROBIOLOGY, Oct. 1991, p. 2821-2828 Vol. 57, No.

10
0099-2240/91/102821-08$02.00/0
Copyright C) 1991, American Society for Microbiology

Propionic Acid Production by a Propionic Acid-Tolerant Strain


of Propionibacterium acidipropionici in Batch
and Semicontinuous Fermentationt
STEVEN A. WOSKOWt AND BONITA A. GLATZ*
Department of Food Science and Human Nutrition, Iowa State University, Ames, Iowa 50011
Received 22 March 1991/Accepted 18 July 1991

A propionic acid-tolerant derivative of Propionibacterium acidipropionici P9 was obtained by serially

Downloaded from http://aem.asm.org/ on October 1, 2018 by guest


transferring strain P9 through broth that contained increasing amounts of propionic acid. After 1 year of
repeated transfers, a strain (designated P200910) capable of growth at higher propionic acid concentrations
than P9 was obtained. An increase in the proportion of cellular straight-chain fatty acids and uncoupling of
propionic acid production from growth were observed for strain P200910. Growth rate, sugar utilization, and
acid production were monitored during batch and semicontinuous fermentations of semidefined medium and
during batch fermentation of whey permeate for both strain P200910 and strain P9. The highest propionic acid
concentration (47 g/liter) was produced by P200910 in a semicontinuous fermentation. Strain P200910
produced a higher ratio of propionic acid to acetic acid, utilized sugar more efficiently, and produced more
propionic acid per gram of biomass than did its parent in all fermentations.

Strains of the genus Propionibacterium are used in several have been used to improve the yield of propionic acid.
industrial processes because of their ability to convert lac- Despite these efforts, the maximum reported yield of propi-
tate and carbohydrates to propionic acid, acetic acid, and onic acid obtained by fermentation is still too low to be
carbon dioxide. The metabolism of glucose by propionibac- economically competitive with chemical synthesis.
teria theoretically yields 2 mol of propionate, 1 mol of The major factor that limits the production of propionic
acetate, and 1 mol of carbon dioxide from 1.5 mol of glucose acid during fermentation is end-product inhibition by the
(29, 41). Propionibacteria are primarily used by the dairy acid (23). To overcome the inhibitory effect of propionic
industry for the production of Swiss-type cheeses. The acid, continuous processes combined with cell recycling
products from the metabolism of lactate are responsible for were used to remove metabolic end products (27, 28).
the characteristic eyes and contribute to the flavor, texture, Although these processes were able to increase the yield, the
and shelf life of Swiss cheese (16). Although they are used production of propionic acid was not high enough to offset
mainly in cheese production, propionibacteria are also used the higher costs of continuous processes. Mutant strains
industrially as silage inoculum (12), as a probiotic agent (24, resistant to end products have been used to increase the
25), and for the production of vitamin B12 (14, 29, 42) and production of ethanol from Clostridium thermocellum (35),
propionic acid (29). butanol from Clostridium acetobutylicum (22), and ethanol
As a preservative, propionic acid extends the shelf life of from yeasts (21).
food products by inhibiting molds and some bacteria (15, 20, To improve the yield of propionic acid, we have developed
23). Although preservatives derived from propionibacterium a propionic acid-tolerant mutant, designated P200910, by
fermentations are available, most propionic acid used by the using a simple enrichment technique. When used in a semi-
food industry is produced by chemical synthesis (29). If continuous fermentation, P200910 produced greater amounts
higher yields of propionic acid could be obtained, production of propionic acid than did the parental strain. We report here
by fermentation might become economically competitive the characterization of P200910 in a batch fermentation and
and might offer several advantages over chemical synthesis. the development of a semicontinuous process. Also, the
These advantages include bacteriocin production (17), which physiological changes that occur in the mutant strain and
can increase the spectrum of antimicrobial activity, the their contribution to acid tolerance are discussed.
ability to label the product as a "natural preservative," and
the opportunity to use food-processing wastes as fermenta-
tion substrates, thus lowering disposal costs. MATERIALS AND METHODS
Several processes have been patented for producing pro-
pionic acid by fermentation (29). Batch methods that use a Strains and culture maintenance. Propionibacterium aci-
variety of substrates typically produce 1 to 3% propionic dipropionici P9 was obtained from the culture collection of
acid in 7 to 14 days (2-4, 29). Other processes, including the Department of Food Science and Human Nutrition, Iowa
fed-batch (16), cell immobilization (6, 7), continuous (2, 4, 6, State University. Strain P9 and its propionic acid-tolerant
8, 29), semicontinuous (13), and multistage (29) processes, derivative P200910 were grown in sodium lactate broth
(NLB) at 32°C (18). Working cultures were maintained on
sodium lactate agar and stored at 4°C. All cultures were
*
Corresponding author. permanently stored at -70°C in NLB supplemented with
t Journal paper no. J-14415 of the Iowa Agriculture and Home 10% glycerol. Fermentation broth (FB) consisted of 0.6%
Economics Experiment Station, Ames (project no. 2826). yeast extract, 0.3% Trypticase, and 3% glucose at pH 7.0.
4: Present address: Great Lakes Biochemical Co., Milwaukee, WI Glucose was sterilized separately in an autoclave and added
51218. aseptically to FB before the start of fermentation.
2821
2822 WOSKOW AND GLATZ APPL. ENVIRON. MICROBIOL.

Culture conditions. Primary cultures were prepared by mark) for pH control. A vessel with 450 ml of FB was
inoculating 10 ml of NLB with isolated colonies from a sterilized in an autoclave for 20 min, and 50 ml of a sterile
sodium lactate agar plate and incubating the broth at 32°C for 30% glucose solution was added aseptically. The tempera-
24 to 36 h. For small-scale cultures, a 1% inoculum of this ture was controlled at 32°C, the agitation rate was 200 rpm,
culture was transferred into 10 ml of the appropriate fresh and the pH was maintained at 7.0 by the addition of 2 N
medium and incubated at 32°C. For 500-ml fermentation NaOH.
experiments, a 1% inoculum of the primary culture was For determination of whether acid production and cell
transferred to 25 ml of FB in a 100-ml Erlenmeyer flask. The growth decreased during a batch fermentation because of
culture was incubated at 32°C and harvested in the exponen- product inhibition or nutrient depletion, spent FB was ob-
tial phase (optical density at 550 nm [OD550], 0.8). All seed tained from a batch fermentation, centrifuged to remove
cultures were incubated without agitation. The entire con- cells, and sterilized by filtration through a 0.22-,um-pore-size
tents of the flask were used to inoculate the fermentor. filter. Samples (9-ml) were placed in screw-cap test tubes,
Growth rate measurements. Small-scale (10-ml) cultures of and 1 ml of filter-sterilized lOx FB was added to each tube.
P9 and P200910 were grown in the desired medium at 32°C, A 2% inoculum of an active 48-h culture of either P9 or
and observed for changes in OD550 with a Spectronic 21 P20091 was added, and the tubes were incubated at 32°C.

Downloaded from http://aem.asm.org/ on October 1, 2018 by guest


spectrophotometer (Milton Roy, Rochester, N.Y.). Specific Culture growth over time was monitored at OD550. Control
growth rates were determined by plotting the natural log (In) tubes contained FB and spent FB with no added nutrients.
of OD550 versus time. Regression analysis was performed on Semicontinuous fermentations. Initial fermentation condi-
the values taken from the linear portion of the curve, and the tions were the same as those described for batch fermenta-
specific growth rates were calculated from the slope of the tions. After the initial 48 h of fermentation, 50 ml of 1Ox FB
least-squares regression line. Analysis of variance was per- and 25 ml of a 60% glucose solution were added while an
formed on data obtained from replicate trials. equal volume of spent FB was removed from the fermentor.
Fatty acid analysis. One-liter cultures of strains P9 and This procedure was repeated every 24 to 36 h until the
P200910 were grown in NLB for 48 h at 32°C to an OD550 of measured amount of propionic acid in FB did not increase.
1.0. The cells were removed by centrifugation and washed Fermentation of whey permeate. Sweet whey was obtained
three times with 0.01 M phosphate buffer (pH 7.0). Fatty from Swiss cheese produced in the Department of Food
acid methyl esters (FAMES) of cellular lipids were prepared Science and Human Nutrition, Iowa State University. The
by the method of Baer et al. (1) with the following modifi- whey was filtered through a DC1OL ultrafiltration unit con-
cations. The methanolic-base reagent was prepared by mix- taining an H5MP01-43 0.1->Lm hollow-fiber filtration car-
ing 170 ml of sodium methoxide (Fluka Chemie AG, Buchs, tridge (Amicon Division, W. R. Grace and Co., Danvers,
Switzerland) with 750 ml of anhydrous methanol. A 0.5-g Mass.) and filtered a second time through an SlOY10 spirally
sample of wet-packed cells, 1 ml of benzene, and 1 ml of wound ultrafiltration cartridge with a 10,000-molecular-
methanolic-base reagent were added to a screw-cap test tube weight cutoff (Amicon). Whey permeate was adjusted to pH
(18 by 150 mm). The tube was sealed and heated in a water 7.0 by the addition of NaOH, filter sterilized, and stored at
bath at 80°C for 20 min. The sample was cooled to room 4°C. Batch fermentations of whey permeate in 9-liter work-
temperature before 3 ml of water and 3 ml of diethyl ether ing volumes were performed in a model NLF 22 fermentor
were added. The contents of the tube were mixed, the lower, (Bioengineering Corp., Wald, Switzerland). The culture was
aqueous phase was removed, and the upper, benzene-ether grown in the fermentor at 32°C and pH 7.0 and agitated at
layer was washed twice with 2 ml of water. Residual water 200 rpm. Temperature, pH, and agitation were monitored
was removed by drying over sodium sulfate crystals. The and controlled by a computer (Nomad System Inc., San
solvent layer, containing the FAMES, was transferred to Jose, Calif.).
3-ml vials and stored under nitrogen at -20°C. Analysis of products. Glucose, lactose, propionic acid, and
Immediately before analysis, the remaining solvent was acetic acid were separated by high-performance liquid chrom-
removed by evaporation under a gentle stream of nitrogen. atography with a model 501 pump (Waters, Division of
The FAMES were resuspended in 50 1.l of hexane, and a 1-,ul Millipore, Milford, Mass.) and an HPX-87H column (Bio-
sample was injected into a model 3700 gas chromatograph Rad, Richmond, Calif.) operated at 65°C with 0.012 N H2SO4
(Varian Aerograph, Palo Alto, Calif.) equipped with a flame (pH 2.0) as the mobile phase at an 0.8-ml/min flow rate.
ionization detector and a fused-silica capillary column (15 m Peaks were detected with a Waters differential refractometer
by 0.32 mm) coated with SP-2330. The column temperature (model R401). Samples for analysis were centrifuged to
was programmed for 100 to 230°C at 4°C/min. The injection remove cells, filtered through 0.22-R,m-pore-size filters, and
and detector temperatures were set at 230°C. An integrator stored at -70°C before analysis. A Maxima 820 software
(model 3396A; Hewlett-Packard, Arondale, Pa.) was used to program (Waters) was used to analyze the data and plot the
analyze data and plot chromatograms. The FAMES were chromatograms. The product concentration was calculated
identified by comparing the retention times with those of by comparing the peak areas with those of external stan-
FAME standards (Foxboro/Analabs, North Haven, Conn.). dards (Aldrich Chemical Co., Milwaukee, Wis.).
The peak heights were measured, and the percentage of each Biomass was determined from a standard curve of optical
peak in the sample was calculated from the ratio of the density versus dry weight. Dry weights for the standard
individual peak height to the sum of the heights of all curve were obtained by filtering culture aliquots through
detected peaks. prerinsed, dried, and weighed 0.22-,um-pore-size filters, rins-
Batch fermentations. Batch fermentations (500-ml working ing the filters with 0.1 M phosphate buffer (pH 7.0), and
volumes) were performed in a model C30 bench-top fermen- drying the filters in a microwave oven (Tappan/O'Keefe &
tor (New Brunswick Scientific Co., New Brunswick, N.J.) Merrit, Chicago, Ill.). The standard curve was plotted from
with an accessory pH controller (model pH-40; New Brun- the mean values of two determinations, and biomass was
swick Scientific Co.) or a Multigen bench-top fermentor obtained from the least-squares regression line.
(New Brunswick Scientific Co.) equipped with a model All fermentations were performed at least twice. Results
TTT2 automatic titrator (Radiometer, Copenhagen, Den- are the averages of these replicate trials.
VOL. 57, 1991 P. ACIDIPROPIONICI PROPIONIC ACID PRODUCTION 2823

0.22
,) 0.20 2
c0
~: 0.16B 3
B
7 9
ZC-a) O0. 14
16
10
a 0.14 8 11
*-'av 0. 12 12
0

A
U 0.06
0)
In 0.04

Downloaded from http://aem.asm.org/ on October 1, 2018 by guest


0.02
0.0 1 0 2 0 3.0 4:0 510 6 0 7:0 60 90so
Propionic Acid Concentration (%) 140 50 160 170 80 so 200 210
FIG. 1. Specific growth rates of P9 (0) and P200910 (0) in NLB
Temperature (c)
containing various amounts of propionic acid. Error bars represent FIG. 2. Typical gas chromatograms of FAMES prepared from
standard deviations for four replicate trials. Specific growth rates strains P200910 (A) and P9 (B).
were determined by plotting the natural log (ln) of OD550 versus
time. Regression analysis was performed on the values taken from
the linear portion of the curve, and the specific growth rates were
calculated from the slope of the least-squares regression line. then inoculating these cultures into NLB with 2% propionic
acid. No difference in growth was observed when inocula
were serially transferred with or without propionic acid.
RESULTS Fatty acid analysis. Fatty acid analysis was performed to
detect any changes in the cellular fatty acid profile associ-
Development of a propionic acid-tolerant strain. The pro- ated with propionic acid tolerance. Typical chromatograms
pionic acid-tolerant strain was derived from strain P9 by are presented in Fig. 2. Table 1 lists the relative amounts of
a modification of the serial dilution method of Lin and the 12 major peaks representing greater than 90% of the total
Blaschek (22). A 1% inoculum of a primary culture of P9 was peaks detected. Eight of the peaks were identified by com-
transferred to a series of tubes with 10 ml of NLB containing paring their retention times with those of known FAME
0.5 to 5% propionic acid. The tubes were incubated at 32°C standards. These peak identifications agreed with previously
for 24 h, and the change in the OD550 was monitored. Cells published fatty acid profiles of propionibacteria (19). The
from the broth with the highest propionic acid concentration predominant fatty acids were 15- and 17-carbon iso- and
that showed growth were repeatedly transferred into fresh anteiso branched-chain and straight-chain fatty acids. Hof-
broth containing that concentration of propionic acid. Once herr et al. (19) also reported the presence of hydrocarbon
the growth rate of the tolerant strain reached approximately peaks tentatively identified as branched-chain 19- and 21-
80% of that of the unchallenged parental strain, the tolerant carbon species. In the present work, peaks 10 and 12 had
strain was transferred into broth containing a slightly higher retention times corresponding to 19:0 and 20:0 straight-chain
amount of propionic acid, and the process was repeated. fatty acids. Four peaks could not be identified. The relative
After 1 year of such repeated transfers, a strain, desig-
nated P200910, that was able to grow at higher concentra-
tions of propionic acid than P9 was obtained. Electron TABLE 1. Identities and relative amounts of the major fatty
micrographs of strain P200910 showed no differences in acids in cellular lipids of strains P9 and P200910, as
morphology from the parental strain (data not shown). Strain determined by gas chromatography
P200910 is identical to the parental strain in Gram reaction, % Fatty acids present in:
fermentation of sucrose, maltose, and mannitol, reduction of Peak Identificationa
nitrate, and pigment production. P9 P200910
Growth characteristics of P200910. A plot of the specific 1 i15:0 16.9 15.4
growth rates for P200910 and P9 in NLB containing different 2 a15:0 18.3 17.5
concentrations of propionic acid is shown in Fig. 1. Analysis 3 nlS:0 13.6 16.6
of variance showed significant differences (P < 0.05) be- 4 HC 6.4 6.1
tween the growth rates of P9 and P200910 at propionic acid 5 HC 2.2 3.5
concentrations between 1 and 7%. No statistically significant 6 i17:0 0.9 0.9
differences were observed between the growth rates of the 7 a17:0 11.4 4.8
two strains at 0 or 8% propionic acid. However, P200910 had 8 n17:0 5.0 8.3
a slightly faster specific growth rate at 8% propionic acid 9 HC 11.4 9.2
10 HCb 8.2 9.6
(0.047 versus 0.033 h-1) and a slightly slower specific growth 11 HC 4.1 3.5
rate (0.185 versus 0.199 h-1) at 0% propionic acid. In NLB, 12 HCC 1.4 4.4
strain P200910 had a longer lag time and grew to a lower final
cell density than did strain P9, even after several transfers. a n, normal; a, anteiso; i, iso; HC, hydrocarbon.
b
The stability of the tolerance of strain P200910 for propi- Unidentified hydrocarbon with a retention time corresponding to 19:0
straight-chain fatty acids.
onic acid was determined by serially transferring this strain c Unidentified hydrocarbon with a retention time corresponding to 20:0
10 times in NLB or in NLB with 0.5% propionic acid and straight-chain fatty acids.
2824 WOSKOW AND GLATZ APPL. ENVIRON. MICROBIOL.

amounts of the predominant fatty acids were different for


40 30
strains P9 and P200910. Strain P200910 had fewer branched-
chain and more straight-chain fatty acids.
36 - 27
Batch fermentations. Data from typical batch fermenta-
tions for strains P9 and P200910 are shown in Fig. 3, and
o 28 - 21
selected parameters are shown in Table 2. The strains had
c~~~~~~~~~~~~~~~~~
24 18
similar lag times (12 h). The growth rate during the exponen-
tial phase was faster for P200910 (0.84 g/liter/h) than for P9
20 - 15 (0.45 g/liter/h), but strain P9 remained in the exponential
16- 12 phase longer (36 h) and reached a higher final biomass. Strain
P9 entered the stationary phase at 48 h; strain P200910 did so
at 28 h.
After 60 h of fermentation, P200910 had produced more
4325 \ ° 243 propionic acid than had P9 per gram of biomass (0.91 versus

Downloaded from http://aem.asm.org/ on October 1, 2018 by guest


o
410
0 0.50). Strain P200910 produced propionic acid faster during
10i 1s 510 515
0
25 30

TTime (h)
35 45 60
the exponential phase and continued production after reach-
- 12 - - 3 ing the stationary phase. Propionic acid production was
fastest between 20 and 38 h for both strains. The final acetic
acid concentration and rate of acetic acid production were
similar for both strains, but more acetic acid was produced
per gram of biomass by strain P200910 than by strain P9 (0.21
versus 0.13). Strain P200910 converted glucose to propionic
FG40 30 acid somewhat more efficiently than did strain P9, but both
strains produced less acid than the theoretical maximum.
3c 27
Propionic acid/acetic acid ratios for both strains were higher
than the theoretical ratio of 2:1.
l21
r n~~~~~~~~~~~~~~~~~~~~
28
Strain P200910 exhibited biomass and propionic acid pro-
24 18s
duction patterns typical of non-growth-associated product
~~~~~~~~~~~~~~~~~15 formation, whereas strain P9 followed typical growth-asso-
V 16 12 an ciated product formation patterns (31). The fastest growth of
strain P200910 occurred earlier thah the period of maximum
V
12 -9 propionic acid production or maximum glucose utilization
(Fig. 4). In contrast, periods of maximum growth and
4 3
glucose utilization coincided for strain P9.
c
0
To determine whether acid and biomass production in
0
a l~1 115 20 30 410 50 55 batch fermentations eventually decreased because of prod-
& Time Ch,) uct inhibition or nutrient depletion, we compared the growth
FIG. 3. Growth, glucose utilization, and acid production by
rates of strains P9 and P200910 in FB, in spent FB containing
strains P200910 (A) and P9 (B) in semidefined miedium in a batch
0.6% propionic acid (SB), and in spent FB supplemented
fermentation process. Symbols: *, biqmiass; K0, glucose; 0, propi- with fresh FB (SB3FB) (Table 3). Both strains showed little
onic acid; 0-, acetic acid. Resuilts are the averages of three replicate growth in SB. Growth rates of P9 and P200910 in SBFB were
trials. 76 and 79%, respectively, of their growth rates in FB.
Nutrient depletion and product inhibition may both play a
role in growth limitation during batch fermentations, al-
though it seems that nutrient depletion has a greater effect.

TABLE 2. Characteristics of fermentations with strains P9 and P20010


Maximum
Maximum
~Final concn Volumetric ye(i)Yieldf (% of
(g/liter) productivity' YP/s5'of:(9/9) theoretical PA/AA
Strain Medium' Processb biomass
production ifpo
(g/liter/h) of:
oo:f:lier/atf:of
maximum) molar
(g/liter) _______ rl
PA AA PA AA X PA A PA AA
P9 FB Batch 23 11.6 3.0 0.33 0.07 0.59 0.30 0.08 54 36 3.0
P200910 FB Batch 15 13.6 3.1 0.39 0.07 0.34 0.35 0.08 64 36 3.5
P9 WP Batch 7.8 8.8 3.4 0.044 0.002 0.37 0.55 0.21 2.1
P200910 WP Batch 1.7 9.2 1.9 0.048 0.007 0.13 0.70 0.14 3.9
P9 FB SCF 78 32 9.1 0.26 0.10 0.74 0.31 0.09 56 41 2.9
P200910 FB SCF 58 47 10.5 0.37 0.12 0.55 0.45 0.10 82 45 3.6
WP, whey permeate.
a

b
Batch, 60-h fermentation in FB or 8-day fermentation in WP; SCF, semicontinuous fermentation for 8 days.
C PA, propionic acid; AA, acetic acid.
d Maximum rate of production calculated from the linear portion of the best-fit curve from Fig. 3, 5, and 6.
e Yield coefficient for the product on a carbon substrate (gram of product produced per gram of sugar utilized). X, biomass.
f Calculated as the percentage of the theoretical maximum yield of 55 g of propionic acid and 22 g of acetic acid per 100 g of glucose (41).
VOL. 57, 1991 P. ACIDIPROPIONICI PROPIONIC ACID PRODUCTION 2825

0.65 0.40 @ 10
C 0.76 0.36 5 9
.. 0.686 0.32 U
v
8
o 0.51 0.28 c u
7
0.24 4)
6 ._40
° 0.42
XS
0.20 L E 5
0
, 0.34 4)
U)
(n
4 0
a 0.25 0.12 , v
3 a0
-
m 0.17 *0.08 ° O
O aL
2
o 0.08 *0.04 0
o 0.00 0.00 0 0

Downloaded from http://aem.asm.org/ on October 1, 2018 by guest


D@ on
Time (h) so
g Time (days)

r 0.70 - 0.40 5
A
b
0. 0.63- -0.36 a5
+j 0.56- 0.32 c>0
10
c 0.49- 0. 28 C0
t- 9
20 0.42- - 0. 24 v
0
1c 8
9 0.35-
0
-
0. 20 °
tl
7
C 0. 2- - 0.16 2 0
4
a 0 .21 -
-l 6
- 0.12 v oi
- 0.14- 4i 5
0.08 2
0 4
0 0.07- 0.04 Cl
-6
J
a 0.00 0.00 0 Q 3
cr i 12 18 24 33 36 42 48 54 60 a

Tiffe Ch) 2
n
n
FIG. 4. Rates of biomass (*) and propionic acid (0) production
by strains P200910 (A) and P9 (B) in semidefined medium in a batch 0
E
fermentation process. Rates were calculated from the averaged 2 Time C(days)
results from three trials. m

FIG. 5. Growth, lactose utilization, and acid production by


strains P200910 (A) and P9 (B) in whey permeate in a batch
Whey permeate fermentations. The performance of P9 and fermentation process. Symbols: *, biomass; O, lactose; 0, propi-
P200910 in batch fertnentations of a natural substrate, cheese onic acid; [, acetic acid. Results are the averages of two replicate
whey permeate, was evaluated (Fig. 5 and Table 2). Growth trials.
and acid production were much slower than in batch fermen-
tations in defined medium. Even after 8 days of fermenta-
tion, only 44 and 36% of the lactose had been used by
P200910 and P9, respectively. Although total biomass and acetic acids per gram of sugar utilized were higher in whey
organic acid concentrations were lower in whey permeate permeate fermentations.
than in semidefined medium, the yields of propionic and Strain P200910 grew to a much lower biomass concentra-
tion than did strain P9, but it produced more propionic acid
per gram of biomass (5.4 versus 1.1). Much more lactose was
converted to propionic acid than to biomass or acetic acid by
TABLE 3. Specific growth rates of P9 and P200910 grown P200910. The propionic acid/acetic acid ratio was about
for 24 h in SB or SBFB double that of strain P9.
Specific growth rate (h-1)a in: Semnicontinuous fermentations. Strains P9 and P200910
Strain were cultivated in semicontinuous fermentations in which
FB SBb SBFBC
fresh FB was added to the vessels at regular intervals, at
P9 0.21 0.05 0.16 which times an equal volume of spent FB was removed.
P200910 0.14 0.04 0.11 Data are presented in Fig. 6 and Table 2. Much higher final
a Specific growth rates were determined by plotting the natural log (In) of concentrations of biomass and organic acids were obtained
OD550 versus time. Regression analysis was performed on the values taken in semicontinuous than in batch fermentations. These were
from the linear portion of the curve, and the specific growth rates were achieved by extending growth and acid production over a
calculated from the slope of the least-squares regression line. Values are longer time and by supplying more sugar as the fermentation
averages from two separate experiments. progressed. Yields of biomass from sugar utilized were
b Obtained from a batch fermentation experiment after 96 h of cultivation;
contained 0.6% propionic acid and no detectable glucose. higher in semicontinuous than in batch fermentations. Yields
I
SB mixed with 1Ox fresh FB at a 9:1 [vol/voll ratio. of acids from sugar utilized were approximately the same as
2826 WOSKOW AND GLATZ APPL. ENVIRON. MICROBIOL.

f-%
L
a#
.6.
=
80 25 25
0
L.E ..j
70 'O 22 22
v
60 u
19 -19 f-
FE
4J 4-i uz
1J 50
0 17 -17
-W C_
ob 10
c 14 -14
40 0
.l
-6 11 11
0 30 u

x 8 8
En .2
20 c
0 6
c
cx
0 0
10 L
CL 3
0
4;

Downloaded from http://aem.asm.org/ on October 1, 2018 by guest


En 0 0
go 0
u
:3
Time Cdays) U
Time th)

f-

45
80 L

40
X 70
35
u 60-
30 ur)
VI)
c
$ 50- 0 v1
25
.1
Nj
D40 u
20

o30 -W 15
4J
c 20-
nO 10
0

0 to 5
L10
0
0 Time (h)
m Time (days)
FIG. 7. Growth, glucose utilization, and acid production by
FIG. 6. Growth and acid production by strains P200910 (A) and strain P200910 during the 36-h periods following the first (A) and
P9 (B) in semidefined medium in a semicontinuous fermentation sixth (B) nutrient additions during semicontinuous fermentation.
process. Symbols: *, biomass; 0, propionic acid; O, acetic acid. Samples were analyzed at 12, 24, and 36 h after nutrient additions.
Results are the averages of three replicate trials. Symbols: *, OD550; O, glucose; 0, propionic acid; E, acetic acid.

concentrations increased through 36 h. After the sixth nutri-


in batch fermentations, with propionic acid production being ent addition, glucose again was exhausted by 24 h, but
favored in semicontinuous fermentations by strain P200910. biomass and organic acid concentrations showed little or no
As was seen in batch fermentations, strain P200910 pro- change.
duced less biomass but more propionic acid than did strain
P9. The growth rates for both strains were rapid through 5 DISCUSSION
days of incubation,.at which time the propionic acid concen-
trations were 26 g/liter for P9 and 38 g/liter for P200910. The The commercial use of fermentation processes for the
amount of propionic acid produced per gram of biomass was production of organic acids is frequently limited by low
higher for P200910 (0.81 g/g) than for P9 (0.41 g/g). Rates of yields and high recovery costs. A strain that is less sensitive
propionic acid production decreased for both strains after 4 to its end products can overcome some of this limitation to
days of incubation. commercial success. Solvent-tolerant strains of bacteria and
Early in the fermentation (days 2 to 4), P9 was able to yeasts have been used to increase the production of ethanol
utilize added glucose within 36 h. After day 4, P9 required at and butanol (21, 22, 35, 36). Lin and Blaschek (22) used a
least 48 h to exhaust the added glucose. After day 6, serial dilution method to increase the butanol tolerance of a
complete glucose utilization did not occur. In contrast, strain of C. acetobutylicum that produced increased
P200910 continued to use glucose at a high rate throughout amounts of butanol. We adapted this procedure to develop
the fermentation. Figure 7 shows the fermentation profile for stable propionic acid tolerance in a Propionibacterium
P200910 in the 36-h period following the first and sixth strain, P200910, that produced more propionic acid but less
nutrient additions. After the first nutrient addition, glucose biomass than its parent.
was exhausted by 24 h, and biomass and organic acid Changes in lipid composition occur in bacteria to assure
VOL. 57, 1991 P. ACIDIPROPIONICI PROPIONIC ACID PRODUCTION 2827

survival under adverse conditions, including exposure to trafiltration cell recycling, with a propionic acid concentra-
organic acids and solvents (26). Warth (39) has shown that tion of 18 g/liter at a maximum volumetric productivity of 2.2
propionic acid-resistant yeast strains are less permeable to g/liter/h. Blanc and Goma (2) and Boyaval and Corre (4) also
propionic acid than are sensitive strains. Baer et al. (1) reported efficient continuous culturing-cell recycling sys-
showed that a butanol-tolerant mutant of C. acetobutylicum tems that produced 17 and 25 g of propionic acid per liter at
responded to the presence of butanol by increasing the maximum volumetric productivities of 5.0 and 14.3 g/liter/h,
percentages of 16:0 and 18:0 fatty acids and decreasing the respectively. The maximum volumetric productivities of
percentages of 16:1 and 18:1 fatty acids in its membrane propionic acid obtained in this study were similar to those
lipids. reported for batch fermentations and for continuous cultur-
Normally, propionibacteria contain predominantly ing systems without cell recycling (3, 5) but were lower than
branched-chain 15:0, 16:0, and 17:0 fatty acids and no those reported for continuous culturing systems with cell
unsaturated fatty acids (19). Gas chromatographic analysis recycling (2, 4, 5).
revealed more straight-chain fatty acids in P200910 than in The 47 g of propionic acid per liter produced in this study
P9. A methyl branch or a cis-double bond in a fatty acid by strain P200910 grown in semidefined medium in a semi-
instead of a saturated straight chain decreases the melting continuous fermentation is 40 to 50% higher than previously

Downloaded from http://aem.asm.org/ on October 1, 2018 by guest


point and increases the surface area of the membrane that reported literature values (2-8, 13, 29, 34) but is less than is
contains the lipid (11). Therefore, a shift from branched- theoretically possible. Yields of propionic and acetic acids
chain to straight-chain fatty acids would likely result in a less and their molar ratios often deviate from theoretical values
fluid, less permeable membrane. This seems to be the derived from known pathways of glucose metabolism. Val-
situation with strain P200910. It would be interesting to ues obtained can be affected by the production of succinate
determine whether there are differences in the flux of propi- (9, 32) or an extracellular polysaccharide (10). During fer-
onate ions across the cell membrane in strains P9 and mentations, we observed both the accumulation of succinate
P200910. in the medium (data not shown) and an increase in viscosity.
Reductions in the yields of cells grown in the presence of The semicontinuous process used in this study shows
a preservative have been reported (30, 37, 38, 40). We promise as a means to produce propionic acid with P200910.
observed that P200910 had a lower biomass yield than had its It was run for 8 days but could be performed in a shorter
parent, yet it continued to utilize glucose in the presence of period. Nutrients were added every 24 to 36 h, but P200910
high propionic acid concentrations. These observations sug- exhausted the glucose within 12 to 16 h. A commercial
gest that this acid-tolerant strain may expend energy to rid process might be run by initially operating the fermentor in a
itself of excess acid. fed-batch mode until the maximum propionic acid concen-
Strain P200910 produced more propionic acid than did its tration is achieved. From that point, spent medium could be
parent, at the expense of acetic acid and biomass. It was thus removed as fresh nutrients are added. The addition of
superior to its parent in the yield of propionic acid from nutrients could be triggered by the glucose level reaching a
substrate, the production of acid per gram of biomass, predetermined lower limit. This process could yield a con-
propionic acid/acetic acid ratio, and percentage of theoreti- tinuous supply of broth with higher propionic acid concen-
cal maximum yield attained. It was able to continue acid trations than have been reported previously.
synthesis as high acid concentrations accumulated in the
medium during fermentation. Such a shift from growth- ACKNOWLEDGMENTS
associated to non-growth-associated acid production may be
advantageous to fermentation productivity, as incoming This research was funded in part by the Iowa Corn Promotion
glucose can be converted principally to organic acids. A Board and the Iowa State University Center for Crops Utilization
Research. S.A.W. held an Iowa State University Biotechnology
decrease in the amount of biomass produced may reduce Fellowship.
downstream processing costs during large-scale production
by reducing the time and energy required for removing REFERENCES
biomass. 1. Baer, S. H., H. P. Blaschek, and T. L. Smith. 1987. Effect of
Whey fermented by propionibacteria is considered to be a butanol challenge and temperature on lipid composition and
natural preservative and is produced commercially for use in membrane fluidity of butanol-tolerant Clostridium acetobutyli-
bakery products (2, 4, 7, 33). Whey permeate is a cheap, cum. Appl. Environ. Microbiol. 53:2854-2861.
readily available substrate for the production of propionic 2. Blanc, P., and G. Goma. 1989. Propionic acid and biomass
acid, but fermentations generally require 10 to 14 days to production using continuous ultrafiltration fermentation of
complete and produce 2 to 11 g of acid per liter (2, 4, 7, 29, whey. Biotechnol. Lett. 11:189-194.
33). We observed that strain P200910 produced more propi- 3. Border, P. M., M. P. J. Kierstan, and G. S. Plastow. 1987.
onic acid from whey permeate than did its parent, but Production of propionic acid by mixed bacterial fermentation.
Biotechnol. Lett. 9:843-848.
fermentation time was too long to make this an economically 4. Boyaval, P., and C. Corre. 1987. Continuous fermentation of
feasible process. A faster-growing derivative of P200910 or sweet whey permeate for propionic acid in a CSTR with UF
an acid-tolerant derivative of a strain fermenting lactose recycle. Biotechnol. Lett. 9:801-806.
rapidly is needed. 5. Carrondo, M. J. T., J. P. S. G. Crespo, and M. J. Moura. 1987.
Several fed-batch, semicontinuous, and continuous propi- Production of propionic acid using a xylose utilizing Propioni-
onic acid fermentation processes have been patented (29). bacterium. Appl. Biochem. Biotechnol. 17:295-312.
These yield 20 to 30 g of propionic acid per liter and require 6. Cavin, J. F., C. Saint, and C. Davies. 1985. Continuous produc-
5 to 14 days to complete (3, 29). Recent work on propionic tion of Emmental cheese flavours and propionic acid starters by
acid production has focused on using continuous methods to immobilized cells of a propionic acid bacterium. Biotechnol.
Lett. 7:821-826.
increase product yields during fermentation. Clausen and 7. Champagne, C. P., C. Baillargeon-Cote, and J. Goulet. 1989.
Gaddy (8) produced 20 g of propionic acid per liter in an Whey fermentation by immobilized cells of Propionibacterium
immobilized-cell reactor. Carrondo et al. (5) reported great- shermanii. J. Appl. Bacteriol. 66:175-184.
est efficiency in a continuous stirred-tank reactor with ul- 8. Clausen, E. C., and J. L. Gaddy. 1984. Organic acids by
2828 WOSKOW AND GLATZ APPL. ENVIRON. MICROBIOL.

continuous fermentation. Chem. Eng. Prog. 80:59-63. 26. McElhaney, R. E. 1976. The biological significance of alterations
9. Crow, V. L. 1987. Citrate cycle intermediates in the metabolism in the fatty acid composition of microbial membrane lipids in
of aspartate and lactate by Propionibacterium freudenreichii response to changes in the environmental temperature, p.
subsp. shermanii. Appl. Environ. Microbiol. 53:2600-2602. 255-281. In M. R. Heinrich (ed.), Extreme environments:
10. Crow, V. L. 1988. Polysaccharide production by propionibacte- mechanisms of microbial adaption. Academic Press, New York.
ria during lactose fermentation. Appl. Environ. Microbiol. 54: 27. Nanba, A., R. Nukada, and S. Nagai. 1983. Inhibition by acetic
1892-1895. and propionic acids of the growth of Propionibacterium sher-
11. de Mendoza, D., and R. N. Farias. 1988. Effect of fatty acid manii. J. Ferment. Technol. 61:551-556.
supplementation on membrane fluidity in microorganisms, p. 28. Neronova, N. M., S. I. Ibragimova, and N. D. Ierusalimskii.
119-148. In R. C. Aloia, C. C. Curtain, and L. M. Gordon (ed.), 1967. Effect of the propionate concentration on the specific
Advances in membrane fluidity, vol. 3. Alan R. Liss, Inc., New growth rate of Propionibacterium shermanii. Mikrobiologiya
York. 36:404-409.
12. Flores-Galarza, R. A., B. A. Glatz, C. J. Bern, and L. D. Van 29. Playne, M. J. 1985. Propionic and butyric acids, p. 731-759. In
Fossen. 1985. Preservation of high moisture corn by microbial M. Moo-Young (ed.), Comprehensive biotechnology, vol. 3.
fermentation. J. Food Prot. 48:407-411. Pergamon Press, Oxford.
13. Fortress, F., and B. B. White. 1954. Propionic acid from wood 30. Salmond, C. V., R. G. Kroll, and I. R. Booth. 1984. The effect of

Downloaded from http://aem.asm.org/ on October 1, 2018 by guest


pulp waste liquor. U.S. patent 2,689,817. Chem. Abstr. 48: food preservatives on pH homeostasis in Escherichia coli. J.
14110. Gen. Microbiol. 130:2845-2850.
14. Hatanaka, H., E. Wang, M. Taniguchi, S. ijima, and T. 31. Sinclair, C. G., and D. Cantero. 1989. Fermentation modelling,
Kobayashi. 1988. Production of vitamin B12 by a fermentor with p. 65-112. In B. McNeal and L. M. Harvey (ed.), Fermentation:
hollow-fiber module. Appl. Microbiol. Biotechnol. 27:470-473. a practical approach. IRL Press, Oxford.
15. Herting, D. C., and E. E. Drury. 1974. Antifungal activity of 32. Smart, J. B., and G. G. Pritchard. 1982. Control of pyruvate
volatile fatty acids on grains. Cereal Chem. 51:74-83. kinase activity during glycolysis and gluconeogenesis in Propi-
16. Hettinga, D. H., and G. W. Reinbold. 1972. The propionic-acid onibacterium shermanii. J. Gen. Microbiol. 128:167-176.
bacteria-a review. II. Metabolism. J. Milk Food Technol. 33. Sobczak, E., and E. Konieczna. 1987. Possibilities of using
35:358-372. propionic bacteria in the production of acetic acid from whey.
17. Hettinga, D. H., and G. W. Reinbold. 1972. The propionic-acid Acta Aliment. Pol. 13:351-357.
bacteria-a review. III. Miscellaneous metabolic activities. J. 34. Stiles, H. R., and P. W. Wilson. 1933. Production of propionic
Milk Food Technol. 35:436-447. acid. U.S. patent 1,932,755. Chem. Abstr. 28:565.
18. Hofherr, L. A., and B. A. Glatz. 1983. Mutagenesis of strains of 35. Tailliez, P., H. Girard, J. Millet, and P. Beguin. 1989. Enhanced
Propionibacterium to produce cold-sensitive mutants. J. Dairy cellulose fermentation by an asporogenous and ethanol-tolerant
Sci. 66:2482-2487. mutant of Clostridium thermocellum. Appl. Environ. Microbiol.
19. Hofherr, L. A., E. G. Hammond, B. A. Glatz, and P. F. Ross. 55:207-211.
1983. Relation of growth temperature to fatty acid composition 36. Van Uden, N. 1985. Ethanol toxicity and ethanol tolerance in
of Propionibacterium strains. J. Dairy Sci. 66:1622-1629. yeasts. Annu. Rep. Ferment. Proc. 8:11-58.
20. Huitson, J. J. 1968. Cereals preservation with propionic acid. 37. Warth, A. D. 1986. Preservative resistance of Zygosaccharo-
Process Biochem. 3:31-32. myces bailii and other yeasts. Aust. CSIRO Food Res. Q.
21. Jimenez, J., and T. Benitez. 1988. Selection of ethanol-tolerant 46:1-8.
yeast hybrids in pH-regulated continuous culture. Appl. Envi- 38. Warth, A. D. 1988. Effect of benzoic acid on growth yields of
ron. Microbiol. 54:917-922. yeasts differing in their resistance to preservatives. Appl. Envi-
22. Lin, Y. L., and H. P. Blaschek. 1983. Butanol production by a ron. Microbiol. 54:2091-2095.
butanol-tolerant strain of Clostridium acetobutylicum in ex- 39. Warth, A. D. 1989. Relationship among cell size, membrane
truded corn broth. Appl. Environ. Microbiol. 45:966-973. permeability, and preservative resistance in yeast species.
23. Lueck, E. 1980. Propionic acid, p. 175-182. In E. Lueck (ed.), Appl. Environ. Microbiol. 55:2995-2999.
Microbial food additives: characteristics, uses, effects. Spring- 40. Warth, A. D. 1989. Transport of benzoic and propionic acids by
er-Verlag, New York. Zygosaccharomyces bailii. J. Gen. Microbiol. 135:1383-1390.
24. Mantere-Alhonen, S. 1983. On the survival of a Propionibacte- 41. Wood, H. G. 1981. Metabolic cycles in the fermentation by
rium freudenreichii culture during in vitro gastric digestion. propionic acid bacteria. Curr. Top. Cell. Regul. 18:255-287.
Meijeritiet. Aikak. 41:19-23. 42. Youngsmith, B., and P. Apiraktivongse. 1983. Vitamin B12
25. Mantere-Alhonen, S. 1987. A new type of sour milk with production from soybean curd whey with Propionibacterium
propionibacteria. Meijeritiet. Aikak. 45:49-61. freudenreichii. J. Ferment. Technol. 61:105-107.

Das könnte Ihnen auch gefallen