Vous êtes sur la page 1sur 13

Enhanced Propionic Acid Fermentation

by Propionibacterium acidipropionici
Mutant Obtained by Adaptation in
a Fibrous-Bed Bioreactor

Supaporn Suwannakham, Shang-Tian Yang

Department of Chemical and Biomolecular Engineering, The Ohio State University,


140 West 19th Avenue, Columbus, Ohio 43210; telephone: 614-292-6611;
fax: 614-292-3769; e-mail: Yang.15@osu.edu
Received 15 August 2004; accepted 18 January 2005

Published online 23 June 2005 in Wiley InterScience (www.interscience.wiley.com). DOI: 10.1002/bit.20473

Abstract: Fed-batch fermentations of glucose by P. acid- Keywords: Propionibacterium acidipropionici; propionic


ipropionici ATCC 4875 in free-cell suspension culture and acid fermentation; fibrous-bed bioreactor
immobilized in a fibrous-bed bioreactor (FBB) were
studied. The latter produced a much higher propionic
INTRODUCTION
acid concentration (71.8  0.8 g/L vs. 52.2  1.1 g/L),
indicating enhanced tolerance to propionic acid inhibition Propionibacteria are Gram-positive, nonspore-forming, non-
by cells adapted in the FBB. Compared to the free-cell
motile, anaerobic, rod-shaped bacteria. They are widely used
fermentation, the FBB culture produced 20–59% more
propionate (0.40–0.65  0.02 g/g vs. 0.41  0.02 g/g), 17% in probiotic and cheese industries (Playne, 1985). They also
less acetate (0.10  0.01 g/g vs. 0.12  0.02 g/g), and 50% can be used to produce vitamin B12, tetrapyrrole compounds,
less succinate (0.09  0.02 g/g vs. 0.18  0.03 g/g) from and propionic acid. The latter is an important chemical used
glucose. The higher propionate production in the FBB was in production of cellulose plastics, herbicides, and perfumes.
attributed to mutations in two key enzymes, oxaloacetate
As a strong mold inhibitor, calcium, sodium, and potassium
transcarboxylase and propionyl CoA: succinyl CoA trans-
ferase, leading to the production of propionic acid from salts of propionate are also widely used as food and feed
pyruvate. Both showed higher specific activity and lower preservatives. Currently, almost all propionic acid is
sensitivity to propionic acid inhibition in the mutant than produced by petrochemical processes, at an annual produc-
in the wild type. In contrast, the activity of PEP carbox- tion rate of &400 million pounds in the U.S. Although there
ylase, which converts PEP directly to oxaloacetate and
has been a high interest to produce propionic acid from
leads to the production of succinate from glucose, was
generally lower in the mutant than in the wild type. For biomass via fermentation with propionibacteria, the rela-
phosphotransacetylase and acetate kinase in the acetate tively low propionic acid concentration, yield, and produc-
formation pathway, however, there was no significant tion rate from the fermentation have been the major barriers
difference between the mutant and the wild type. In for economical production of propionic acid from glucose
addition, the mutant had a striking change in its morphol-
ogy. With a threefold increase in its length and &24%
and other fermentable sugars.
decrease in its diameter, the mutant cell had an &10% The problems associated with conventional propionic acid
higher specific surface area that should have made the fermentation largely stem from the strong end product
mutant more efficient in transporting substrates and inhibition caused by propionic acid even at a very low
metabolites across the cell membrane. A slightly lower concentration of 10 g/L (Gu et al., 1998; Hsu and Yang,
membrane-bound ATPase activity found in the mutant
1991). The conventional batch propionic acid fermentation
also indicated that the mutant might have a more efficient
proton pump to allow it to better tolerate propionic acid. In usually takes more than 3 days to reach &20 g/L propionic
addition, the mutant had more longer-chain saturated acid, with a product yield usually less than 0.45 g/g sugar
fatty acids (C17:0) and less unsaturated fatty acids (C18:1), fermented. Attempts to improve propionic acid fermentation
both of which could decrease membrane fluidity and in terms of its yield, final product concentration, and
might have contributed to the increased propionate
production rate have resulted in the development of new
tolerance. The enhanced propionic acid production from
glucose by P. acidipropionici was thus attributed to both a bioprocesses and mutant strains but with limited success
high viable cell density maintained in the reactor and (Emde and Schink, 1990; Huang et al., 1998; Jin and Yang,
favorable mutations resulted from adaptation by cell 1998; Lewis and Yang, 1992; Paik and Glatz, 1994; Rickert
immobilization in the FBB. ß 2005 Wiley Periodicals, Inc. et al., 1998; Solichien et al., 1995; Woskow and Glatz, 1991).
Among all these previous efforts, a fibrous-bed immobilized-
Correspondence to: Shang-Tian Yang
Contract grant sponsors: U.S. Department of Agriculture; Consortium for
cell bioreactor has shown to be a promising technology that
Plant Biotechnology Research, Inc. can significantly improve volumetric productivity, product
Contract grant number: CSREES 99-35504-7800 yield, and final product concentration in several organic acid

ß 2005 Wiley Periodicals, Inc.


fermentations (Huang and Yang, 1998; Lewis and Yang, (Marubishi MD-300) through a recirculation loop (&1 m
1992; Silva and Yang, 1995; Yang et al., 1994, 1995; Zhu and long, tubing ID: 3.1 mm; Microflex Norprene 06402-16; Cole
Yang, 2003; Zhu et al., 2002). Palmer, Chicago, IL) and operated under well-mixed
The potential of using the fibrous bed bioreactor to produce conditions with pH and temperature controls. The FBB
high-concentration propionic acid from glucose was evaluated was constructed by packing a spiral wound cotton towel into
in this study. We found that a final concentration of more than a glass column bioreactor, and had a working volume
70 g/L of propionic acid could be produced from glucose in a of &690 mL. A detailed description of the reactor construc-
fed-batch fermentation. This propionic acid concentration was tion has been given elsewhere (Silva and Yang, 1995). After
not only much higher than that from a similar fed-batch inoculation with &100 mL of cell suspension (OD600 & 2.0)
fermentation with free cells grown in a conventional stirred- into the fermentor, cells were grown for 3–4 days to reach an
tank bioreactor, but also higher than the highest concentration optical density (OD600) of &3.5. The fermentation broth was
(57 g/L) previously reported for an acid-tolerant strain then circulated at a flow rate of &30 mL/min through the
immobilized in calcium alginate beads (Paik and Glatz, FBB to allow cells to attach and be immobilized in the fibrous
1994). To understand the underlying mechanisms contributing matrix. The process continued for 60–72 h until most cells
to the improved propionic acid production and tolerance by had been immobilized in the FBB. The medium circulation
cells immobilized in the FBB, the effects of propionic acid on rate was then increased to &80 mL/min, and the fermenta-
both the original (wild type) and adapted (from the FBB) tion was run at a repeated batch mode to obtain a high cell
cultures were also studied and compared, with respect to their density in the fibrous bed. Fed-batch fermentation with pulse
specific growth rates, cellular activities of several key enzymes additions of concentrated glucose solution was then per-
in the dicarboxylic acid pathway, and membrane-bound formed to study the fermentation kinetics and to evaluate the
ATPase. In addition to significant differences in some of these achievable maximum propionic acid concentration. Samples
factors studied, it was also found that there were significant were taken at regular time intervals throughout the
changes in the membrane fatty acid composition and cell fermentation for analyses. At the end of the fed-batch
morphology. Based on these results, how the adapted (mutant) fermentation, adapted cells (mutants) in the FBB were
cells from the FBB acquired the ability to tolerate and produce removed from the fibrous matrix by vortexing the matrix in
more propionic acid is also discussed in this article. sterile distilled water. The cells were collected for viability
assay and subcultured in serum bottles for further analyses.
MATERIALS AND METHODS

Culture and Media Effect of Propionic Acid on Cell Growth

Propionibacterium acidipropionici ATCC 4875 was grown To determine the inhibition effect of propionic acid on cell
in a synthetic medium containing (per liter): 10 g yeast growth, cells were grown under anaerobic conditions in
extract (Difco Laboratories, Detroit, MI), 5 g Trypticase serum tubes containing 10 mL of the synthetic medium with
(BBL), 0.25 g K2HPO4, 0.05 g MnSO4 and 50–100 g glucose 20 g/L glucose and varying amounts of propionic acid (0–
as the substrate. The medium was sterilized by autoclaving at 100 g/L). Cell growth was followed by measuring the optical
1218C, 15 psig for 30 min. The stock culture was kept in density (OD) at 600 nm. The experiment was done in
serum bottles under anaerobic conditions at 48C. duplicate. The specific growth rates at various initial
propionic acid concentrations were then estimated from the
Free-Cell Fermentation semilogarithmic plots of the OD vs. time.

Unless otherwise noted, the fermentation was carried out in a


5-L fermentor (Marubishi MD-300) containing 2 L of the Enzyme Assays
synthetic medium, at 328C, pH 6.5, and 100 rpm for agitation.
Anaerobiosis was established by sparging the medium with Cells grown in the synthetic medium (50 mL) at 328C to the
N2 for &30 min, and thereafter the fermentor headspace was exponential phase (OD600 & 1.8) were harvested, washed,
maintained under 5 psig N2. The fermentor was inoculated and resuspended in 15 mL of 25 mM Tris/HCl (pH 7.4). The
with &100 mL of a fresh culture grown in a serum bottle cell suspension was then sonicated to break cell walls and
(OD600 & 2.0), and liquid samples were withdrawn at regular centrifuged at 10,000 rpm for 1 h to remove cell debris. The
time intervals. The glucose level in the fermentation broth cell extracts were kept cold on ice before they were used in
was replenished by pulse feeding a concentrated glucose the enzyme activity assays. The protein content of the
solution to allow the fermentation to continue until reaching extracts was determined in triplicate by Bradford protein
the maximum propionic acid concentration. assay (Bio-Rad, Hercules, CA) with bovine serum albumin as
the standard.
Phosphotransacetylase was assayed by the method of
Immobilized-Cell Fermentation
Andersch et al. (1983). The assay solution (final total volume:
The fermentation was carried out with cells immobilized in a 200 mL) contained: 0.1M potassium phosphate buffer
fibrous-bed bioreactor (FBB) connected to a 5-L fermentor (pH 7.4), 0.2 mM acetyl-CoA, 0.08 mM 5,50 -dithio-bis

326 BIOTECHNOLOGY AND BIOENGINEERING, VOL. 91, NO. 3, AUGUST 5, 2005


(2-nitrobenzoate) (DTNB), and the cell extract (120 mL). Membrane-Bound ATPase Assay
The enzyme activity was determined at 328C by measuring
Cells were grown at 328C in the synthetic medium (50 mL)
the absorbance at 405 nm (SpectraMax 250) following the
containing 0.4% (w/v) glycine to the exponential phase
liberation of coenzyme A, which has a molar extinction
(OD600 & 1.8) or stationary phase (OD600 & 3.0), and
coefficient of 13.6 mM1cm1 (Ellman, 1959). One unit of
harvested and suspended in 2 mL of TE buffer (10 mM
activity is defined as the amount of enzyme converting 1 mmol
Tris-HCl, 1 mM EDTA, pH 8.0). Then, mutanolysin
of acetyl-CoA per minute.
(Sigma) at the final concentration of 100 mg/mL was added
Acetate kinase was assayed by the method of Allen et al.
to lyse the cells at 378C for 40 min. The crude membrane
(1964). The assay mixture (final total volume: 300 mL)
fraction was obtained by centrifugation at 48C for 30 min and
contained: 81 mM Tris/HCl buffer (pH 7.4), 4 mM ATP,
was suspended in 2.5 mL of 100 mM PIPES buffer (pH
4 mM MgCl2, 1.6 mM phosphoenolpyruvate, 81 mM
5.95). ATPase activity was determined based on the method
potassium acetate, 0.4 mM NADH, 0.4 U/mL pyruvate
of Gutierrez and Maddox (1992) by measuring the release of
kinase, 0.4 U/mL lactate dehydrogenase, and the cell extract
inorganic phosphate (Pi) from ATP in the reaction mixture
(160 mL). The reaction rate at 258C was followed by
(total volume: 500 mL) containing 5 mM ATP, 10 mM
measuring the absorbance at 340 nm, which decreased
MgCl2 and 50 mL of the crude membrane fraction. The
linearly with time. A control without acetate was used as the
mixture was incubated at 378C for 20 min and the reaction
blank to correct for any ADP that might be present in the ATP
was stopped by adding 1 mL of 15% (w/v) ice-cold
preparation.
trichloroacetic acid. The mixture was then centrifuged twice
Oxaloacetate transcarboxylase was assayed by the method
at 13,200 rpm for 5 min, and the inorganic phosphate Pi in the
of Wood et al. (1969). Briefly, the reaction mixture (final total
supernatant was determined spectrophotometrically at
volume: 300 mL) contained: 0.1 mM NADH, 0.2 mM
830 nm (Shimadzu UV-1601) following the molybdenum
methylmalonyl CoA, 10 mM sodium pyruvate, 350 mM
blue method (Vogel, 1978). One unit of activity is defined as
potassium phosphate (pH 6.8), 2 U/mL malate dehydro-
the amount of enzyme that releases 1 mmol of Pi per minute,
genase, and the cell extract (70 mL). The reaction rate at 258C
and the specific activity of ATPase is defined as the unit of
was followed by measuring the absorbance at 340 nm, which
activity per mg biomass.
decreased linearly with time, for at least 4 min. The reaction
mixture without methylmalonyl CoA was used as the blank.
Propionyl CoA:succinyl CoA transferase (CoA transfer- Cell Membrane Fatty Acid Analysis
ase) was assayed following the method of Schulman and
Cells were grown to the exponential phase and then analyzed
Wood (1975). The reaction mixture (250 mL) consisted of:
for their membrane fatty acid composition with fatty acid
100 mL of Mixture 1 (250 mM Tris/HCl buffer (pH 8.0), 1
methyl ester (FAME) analysis done by Microcheck, Inc.
mM sodium malate, 2.5 mM NAD), 10 mL of 1.5M sodium
(Northfield, VT).
acetate, 10 mL of Mixture 2 (220 units of malate
dehydrogenase, 35 units of citrate synthase, and 0.1M
potassium phosphate buffer (pH 6.8) to make up a 1-mL
Cell Viability Assay
volume), 10 mL of 15 mM succinyl CoA, and 60 mL of the
cell extract. The reaction rate at 258C was followed by Cell viability assay followed the method previously described
measuring the absorbance at 340 nm, which increased by Glenner (1977). Briefly, 2,3,5-triphenyl-2H-tetrazolium
linearly with time, for 3–5 minutes. chloride (final concentration: 1 g/L) was added to the cell
Phosphoenolpyruvate carboxylase (PEP carboxylase) was suspension and incubated at room temperature for 30 min. for
assayed by the method of Maeba and Sanwal (1969). The color development. After centrifugation at 10,000 rpm for
assay mixture, with a total volume of 300 mL, contained: 10 min, the cell pellets were collected and then suspended in
66.67 mM Tris-HCl buffer (pH 9.0), 66.67 mM NADH, 10 methanol to extract the pink color. The absorbance of the
mM MgCl2, 10 mM NaHCO3 (freshly prepared), 3.33 mM methanol extract at 485 nm, which is proportional to the
PEP, 6.25 U/mL malate dehydrogenase, and the cell extract viable cell number, was measured with a spectrophotometer
(180 mL). The reaction rate at 258C was followed by (Shimadzu, UV-1601). The cell viability (%) is reported with
measuring the absorbance at 340 nm. A blank control had no cells cultured in serum bottles and harvested in the
PEP in the reaction solution. exponential phase as the control with 100% viability.
Unless otherwise noted, the activities of above enzymes
were determined on the basis of the molar extinction
Scanning Electron Microscopy
coefficient of 6.2 mM1cm1 for NADH (Van der Werf
et al., 1997) and one unit of activity is defined as the amount Cells were laid by gentle filtration of cell suspension onto the
of the enzyme catalyzing the reaction at a rate of producing or filter paper (Fisherbrand Grade P5, Fisher Scientific,
consuming 1 mmole of NADH per minute. The specific Hampton, NH). The samples were fixed in 2.5% (w/v)
enzyme activity is defined as the unit of enzyme activity per glutaraldehyde for 15 h at 48C and rinsed with distilled water
mg of total protein. All enzyme activity assays were done in twice. The samples were processed through a progressive
duplicate. dehydration with 20–100% ethanol at 10% increment, dried

SUWANNAKHAM AND YANG: ENHANCED PROPIONIC ACID FERMENTATION BY P. ACIDIPROPIONICI MUTANT 327
with HMDS, and coated with gold/palladium. The samples controller, a pump (LC-10Ai), an automatic injector (SIL-
were scanned and photographed with a Philips XL 30 10Ai), a column oven (CTO-10A), a refractive index detector
scanning electron microscope at 15 kV. (RID-10A), and a computer with data analysis software
(Shimadzu version 4.2).
Analytical Methods
RESULTS AND DISCUSSION
Fermentation samples were analyzed for the optical density
at 600 nm (OD600) using a spectrophotometer (Sequoia-
Fermentation Kinetics
turner, Model 340). One unit of OD600 was equivalent to
0.435 g/L of cell dry weight. The concentrations of glucose Figure 1 shows typical kinetics for fed-batch fermentations
and acid products (mainly, propionic, acetic, and succinic with free cells in a stirred-tank fermentor and immobilized
acids) were analyzed by high-performance liquid chromato- cells in the fibrous bed bioreactor. These fed-batch
graphy (HPLC) with a Bio-Rad HPX-87H organic acid fermentations were allowed to continue until cells ceased
analysis column at 458C using 0.01 N H2SO4 at 0.6 mL/min to consume glucose or produce propionic acid. The purpose
as the eluent. The HPLC system (Shimadzu) consisted of a of the experiment was to allow cells to gradually adapt to the

Figure 1. Fed-batch fermentations of glucose by P. acidipropionici at pH 6.5, 328C. (A) Free-cell fermentation. (B) Immobilized-cell fermentation in the
FBB.

328 BIOTECHNOLOGY AND BIOENGINEERING, VOL. 91, NO. 3, AUGUST 5, 2005


high-propionate concentration environment to test the third of the total liquid volume. It should be noted that the
maximum propionic acid concentration that can be produced specific productivity in the immobilized-cell fermentation
by the bacteria. As seen in Figure 1, the immobilized-cell was approximately 80% of that in the free-cell fermentation
fermentation in the FBB produced more propionate from because of the much higher cell density in the immobilized-
glucose and reached a higher final propionic acid concentra- cell system (11.5 g viable cell/L vs. 2.4 g viable cell/L for the
tion of 71.8 g/L, which was &40% higher than that from the free-cell fermentation). A lower specific productivity was
free-cell fermentation (52.2 g/L). It should be noted that 71.8 often found in immobilized-cell fermentations (Goncalves
g/L is the highest propionic acid concentration ever produced et al., 1992; Krischke et al., 1991; Norton et al., 1994;
in propionic acid fermentation. Previously, the maximum Senthuran et al., 1997) because of nutrient limitation and
propionate concentration produced from glucose was 57 g/L reduced cell growth.
by a propionate-tolerant strain P. acidipropionici P200910 Compared to the free-cell fermentation, the immobilized-
immobilized in calcium alginate beads (Paik and Glatz, cell fermentation in the FBB produced more propionic acid
1994). A propionic acid concentration of 65 g/L was also and less succinic and acetic acids from glucose. Figure 3
attained from fermentation with cells immobilized in the shows cumulative consumption of glucose and production of
FBB using de-lactose whey permeate as the substrate (Yang propionate, acetate, and succinate in the fed-batch fermenta-
et al., 1995). The higher final product concentration obtained tions. It is noted that the overall propionic acid yield from the
in the fermentation would facilitate product separation and immobilized-cell fermentation was high, &0.65 g/g, for
recovery, and significantly reduce the production costs (Van propionic acid concentrations up to 45 g/L, but then
Hoek et al., 2003). decreased to 0.4 g/g as the propionic acid concentration
The higher propionic acid concentration produced in the increased to 71 g/L. In contrast, the average propionic acid
immobilized-cell fermentation indicated that cells in the yield from the free-cell fermentation was 0.41 g/g. The lower
FBB were possibly less sensitive to propionic acid inhibition. propionic acid yield in the free-cell fermentation can be
As shown in Figure 2, at comparable propionic acid partially attributed to the fact that more carbon sources are
concentrations, the volumetric productivity for propionic used for cell growth and energy production. In the FBB with
acid was much higher in the immobilized-cell fermentation high cell density, cell growth was limited (Huang et al., 2002;
than in the free-cell fermentation. The higher volumetric Yang et al., 1994; Zhu and Yang, 2003), as indicated by the
productivity in the immobilized-cell fermentation could also relatively low OD value throughout the fermentation (see
be attributed to the high cell density (>45 g/L reactor Fig. 1). Furthermore, more acetate and succinate were
volume) and cell viability in the FBB. At the end of the fed- produced in the free-cell fermentation than in the immobi-
batch fermentation, the total amount of cells in the FBB lized-cell fermentation. As can be seen in Table I, the free-
was &32 grams, of which more than 95% was immobilized cell fermentation produced 20% more acetate (0.12 g/g vs.
in the fibrous matrix, with an average cell viability of greater 0.10 g/g) and 100% more succinate (0.18 g/g vs. 0.09 g/g)
than 70%. The volumetric productivities shown in Figure 2 than the immobilized-cell fermentation. The reduced cell
were calculated based on the total volume of the liquid growth in the FBB would lower the ATP demand for biomass
medium (2 L) in the fermentation, instead of the actual formation and thus result in decreased acetate production as
working volume of the FBB (0.69 L), which was about one the acetate formation pathway is the major pathway for ATP
production in propioinibacteria (Goswami and Srivastava,
2000; Rickert et al., 1998). The lower succinate production in
the immobilized-cell fermentation can be attributed to
changes in the activities of several key enzymes in the
dicarboxylic acid pathway, which will be discussed later in
this article.
The results from the fermentation experiments clearly
indicate that cells immobilized in the FBB acquired an ability
to tolerate and produce a higher propionate concentration that
could not be achieved by the original culture grown in free
suspension. Also, there was a significant shift in the
metabolic pathway, leading to a higher propionate yield
from glucose and decreased production of acetate and
succinate. Further experiments were thus conducted to study
the underlying changes or causes that allowed the adapted
culture in the FBB to produce more propionate at a higher
concentration than previously reported. To determine if there
were any phenotypic changes in the FBB culture, cells in the
Figure 2. Effect of propionic acid on the volumetric productivity of
FBB were removed and grown as suspension culture to test
propionic acid in fed-batch fermentations with free cells and immobilized their propionate tolerance. The activities of several key
cells. enzymes in the acid-forming pathways and cell membrane

SUWANNAKHAM AND YANG: ENHANCED PROPIONIC ACID FERMENTATION BY P. ACIDIPROPIONICI MUTANT 329
FBB at various initial concentrations of propionic acid in the
growth media. Although the wild type had a slightly higher
growth rate in the absence of propionic acid, the adapted
culture had a significantly higher growth rate for the
propionate concentrations tested between 5 g/L and 60 g/L.
Similar results have been reported for a propionate-tolerant P.
acidipropionici strain, which also had a slightly lower
specific growth rate than the wild-type strain (0.185 h1 vs.
0.199 h1) in the absence of propionic acid but a higher
growth rate at 8% propionic acid (0.047 h1 vs. 0.033 h1)
(Woskow and Glatz, 1991). In this study, the specific growth
rate reduced by 70% for the wild type and 50% for the FBB
adapted culture, respectively, as the propionate concentration
increased to 10 g/L. There was minimal growth when the
propionate concentration was 80 g/L and higher.
The effect of propionic acid on cell growth can be
described by the following noncompetitive product inhibi-
tion model:

mmax Ki 1 1 1
m¼ or ¼ þ P
Ki þ P m mmax mmax Ki

where m is the specific growth rate (h1), max is the


maximum specific growth rate (h1), K i is the inhibition rate
constant (g/L), and P is the propionic acid concentration
(g/L). The values of max and Ki were determined from the
linear plot of 1/m vs. P shown in the inset of Figure 4, and are
given in Table II. Compared to the wild type, the adapted
culture had a slightly lower mmax (0.133  0.008 vs.
0.154  0.017 h1), but a much higher Ki (8.93  0.77 vs.
4.33  0.67 g/L). It is clear that the adapted culture is less
sensitive to propionic acid inhibition.

Acid-Forming Enzyme Activities


The cellular activities of several key enzymes in the
dicarboxylic acid pathway (see Fig. 5) that would have
major effects on the production of propionic acid, acetic acid,
and succinic acid were examined and compared between the
mutant strain and the wild type. As shown in Figure 6, the
Figure 3. Comparison of product yields from glucose in fed-batch activities of both oxaloacetate transcarboxylase and propio-
fermentations with free cells and immobilized cells of P. acidipropionici. nyl CoA:succinyl CoA transferase (CoA transferase), two
(A) Cumulative propionic acid production vs. glucose consumption. (B) key enzymes in the pathway leading to propionic acid
Cumulative acetic acid production vs. glucose consumption. (C) Cumulative production, were significantly higher in the mutant than in the
succinic acid production vs. glucose consumption. The product yields can be
estimated from the slopes of these plots.
wild type. Oxaloacetate transcarboxylase, a key enzyme in
the pathway for propionate synthesis, catalyzes a coupled
ATPase as well as membrane fatty acid composition were reaction of pyruvate to oxaloacetate and methylmalonyl CoA
also examined and compared with those of the wild type to propionyl CoA. It is responsible for supplying the carbon
culture used in seeding the bioreactor. flux from the central carbon metabolism toward the end
product, propionate. CoA transferase also catalyzes a
coupled reaction of succinate to succinyl CoA and propionyl
Propionic Acid Inhibition
CoA to propionate. It was found that CoA transferase was
Propionic acid is a strong growth inhibitor (Balamurugan less active but more sensitive to propionic acid inhibition
et al., 1999) and 1% (w/v) propionic acid in the medium could than was transcarboxylase for both the mutant and wild type.
reduce the specific growth rate of propionibacteria by more Clearly, the reaction catalyzed by CoA transferase was the
than 50% (Jin and Yang, 1998). Figure 4 compares the specific rate-limiting step in the pathway leading to propionic acid
growth rates for the wild type and the adapted culture from the production in P. acidipropionici.

330 BIOTECHNOLOGY AND BIOENGINEERING, VOL. 91, NO. 3, AUGUST 5, 2005


Table I. Comparison of product yields and maximum propionic acid concentrations from fed-batch
fermentations with free cells and immobilized cells in the FBB.

Free-cell fermentations
Immobilized-cell
Wild type Mutanta fermentation

Max. propionic acid concentration (g/L) 52.2  1.1 51.5  0.9 71.8  0.8
Product yield (g/g)
Propionic acid 0.41  0.02 0.47  0.01 0.40  0.65  0.02b
Acetic acid 0.12  0.02 0.11  0.01 0.10  0.01
Succinic acid 0.18  0.03 0.09  0.01 0.09  0.02
Total viable cell concentration (g/L) 2.41 5.67 11.48c
Total carbon recovery (%)d 88.2  8.5 87.1  3.7 95.7  6.0

Note: Mean  standard deviation.


a
Cells from the FBB after adaptation to high propionic acid concentration in the fed-batch fermentation.
b
Yield was &0.65 for propionic acid concentrations up to &45 g/L and then decreased with increasing
propionic acid concentration to &0.40 at &71 g/L.
c
Including all viable cells in suspension and immobilized in the fibrous matrix; based on the total liquid volume
of 2 L.
d
Based on the assumption that 46.2% of the biomass was carbon. The amount of CO2 produced was estimated
based on the assumption that one mole of CO2 was coproduced with each mole of acetic acid in the fermentation.

On the other hand, the activities of phosphotransacetylase 60 g/L of propionic acid. The increased PTA activity at
(PTA) and acetate kinase (AK) in the mutant were either high propionic acid concentrations could be an induced cell
about the same as or slightly lower than those of the wild type response to direct more pyruvate towards the acetate-forming
(Fig. 7). PTA and AK are two key enzymes in the pathway pathway instead of the propionate-forming pathway.
from pyruvate to acetate, which is the major supply route Also, the increased PTA activity would increase the cellular
for ATP. PTA catalyzes the reaction of acetyl CoA to acetyl level of acetyl phosphate, which can function as a global
phosphate, and AK catalyzes the conversion of acetyl regulator in a metabolic network (McCleary et al., 1993).
phosphate to acetate. Both PTA and AK were highly sensitive Acetyl phosphate can be used in the phosphorylation of a
to propionic acid inhibition, but PTA had a much group of proteins regulating cell’s responses to environ-
lower activity and appeared to be the rate-limiting enzyme mental stimulations. The global phosphorylation could lead
in the acetate-formation pathway. However, it was not to acetate formation from acetyl phosphate (Wanner and
clear why the activity of PTA increased with increasing Wilmes-Riesenberg, 1992) without the activity of acetate
the propionic acid concentration from the minimum level at kinase.

Figure 4. Effects of propionic acid on specific growth rates of P. acidipropionici wild type and mutant from the FBB. Inset shows the determination of rate
constants in the noncompetitive inhibition of propionic acid. The curves from the model predictions simulate the data (symbols) well.

SUWANNAKHAM AND YANG: ENHANCED PROPIONIC ACID FERMENTATION BY P. ACIDIPROPIONICI MUTANT 331
Table II. Comparison of rate constants for specific growth rate and several key enzymes in P: acidipropionici wild type and mutant from the FBB.

Wild type Mutant from FBB

mmax (h1) or vmax (U/mg) Ki (g/L) mmax (h1) or vmax (U/mg) Ki (g/L)

Specific growth rate 0.154  0.017 4.33  0.67 0.133  0.008 8.93  0.77
OAA transcarboxylase 0.111  0.001 84.8  4.9 0.138  0.014 451.9  42.9
CoA transferase 0.027  0.001 10.1  0.2 0.059  0.004 13.5  0.5
Phosphotransacetylase 0.0012  0.0001 18.4  0.5 0.0011  0.0001 19.0  0.8
Acetate kinase 0.019  0.001 10.2  0.8 0.021  0.002 10.6  0.5

Note: Mean  standard deviation; n ¼ 2. Both wild type and mutant cells were grown in suspension cultures for specific growth rate determination and
enzyme activity assays.

PEP carboxylase is the enzyme that catalyzes the reaction In summary, the enzyme activity assay results were in good
from PEP directly to oxaloacetate, thus leading to the accordance with the fermentation results that the mutant
production of succinate from glucose. As shown in Figure 8, produced more propionic acid and less acetic and succinic
the activity of PEP carboxylase was generally lower in the acids than did the wild type. It is clear that changes in various
mutant than in the wild type. Furthermore, unlike other acid- key enzyme activities in the dicarboxylic acid pathway led to
forming enzymes, PEP carboxylase was relatively stable in a metabolic shift to favor more propionic acid production
the presence of propionic acid up to &80 g/L. In fact, the over acetic and succinic acids. These results also provide
enzyme activity in the wild type seemed to increase insights on how one could metabolically engineer the
significantly when the propionic acid concentration increased propionibacterium to further increase propionic acid produc-
from 10 to 60 g/L. These characteristics could explain tion and reduce byproduct formations.
why succinate production increased significantly later in the Propionic acid is an inhibitor to oxaloacetate transcarbox-
fed-batch fermentations when the propionic acid concentra- ylase, CoA transferase, PTA, and AK (Figs. 6 and 7), and the
tion was higher (see Fig. 3C) and why the adapted culture or inhibition can be modeled by the following non-competitive
mutant in the FBB had much lower succinic acid production inhibition kinetics equation:
as compared to the wild type. This finding is also consistent
with the result obtained in an extractive fermentation vmax Ki 1 1 1
v¼ or ¼ þ P
where propionic acid was continuously removed from Ki þ P v vmax vmax Ki
the fermentor and the production of succinic acid was
reduced to almost zero as a result of the low propionic acid where v is the specific activity of enzyme (U/mg), vmax is the
concentration maintained in the fermentation broth (Jin and maximum specific activity (U/mg), Ki is the inhibition rate
Yang, 1998). constant (g/L), and P is the propionic acid concentration

Figure 5. The dicarboxylic acid pathway for the conversion of glucose to propionic, acetic, and succinic acids. The five key enzymes assayed in this study are
labeled in the pathway.

332 BIOTECHNOLOGY AND BIOENGINEERING, VOL. 91, NO. 3, AUGUST 5, 2005


Figure 6. Effect of propionic acid on oxaloacetate transcarboxylase and Figure 7. Effect of propionic acid on phosphotransacetylase and acetate
propionyl CoA:succinyl CoA transferase activities of P. acidipropionici wild kinase activities of P. acidipropionici wild type and mutant from the FBB.
type and mutant from the FBB. The curves from the non-competitive The curves show the noncompetitive inhibition model predictions
inhibition model simulate the data (symbols) well. (A) Oxaloacetate trans- that simulate the data (symbols) well for propionic acid concentrations up
carboxylase; (B) CoA transferase. to 60 g/L. (A) Phosphotransacetylase; (B) Acetate kinase.

propionic acid concentration and yield obtained in the fed-


batch fermentation by cells immobilized in the FBB.
(g/L). The best values of vmax and Ki for these enzymes were
determined from nonlinear regression of the data and are Membrane-Bound ATPase
listed in Table II. The higher vmax indicates a more active The uncoupling effect of propionic acid on oxidative
enzyme, whereas the higher Ki indicates that the enzyme is phosphorylation may interfere with the establishment and
less sensitive to propionic acid inhibition. Compared to the
wild type, oxaloacetate transcarboxylase and CoA transfer-
ase in the mutant not only were more active, but also less
sensitive to propionic acid inhibition, especially when the
propionic acid concentration was lower than 60 g/L. In fact,
the mutant’s transcarboxylase activity was almost unaffected
until the propionic acid concentration was higher than 60 g/L
(Fig. 6A). In contrast, both PTA and AK in the mutant and
wild type had almost the same vmax and Ki , indicating their
similar activity and sensitivity to propionic acid inhibition.
Compared to the wild type, the higher propionate tolerance of
the mutant was mainly attributed to the improvements in its
oxaloacetate transcarboxylase and CoA transferase that
could maintain relatively high activities and thus allow the
cells to survive even at elevated propionate concentrations.
The alteration in the sensitivity to propionic acid inhibition of Figure 8. Effect of propionic acid on PEP carboxylase activity of P.
these two enzymes also contributed to the increased final acidipropionici wild type and mutant from the FBB.

SUWANNAKHAM AND YANG: ENHANCED PROPIONIC ACID FERMENTATION BY P. ACIDIPROPIONICI MUTANT 333
maintenance of a functional pH-gradient across the cell strains, but the former appeared to have a consistently lower
membrane for the transport of metabolites (Gutierrez and ATPase activity level, an indication that the mutant was more
Maddox, 1992). To maintain a proper pH gradient, the extra efficient in pumping out protons and could survive in a higher
protons must be pumped out at the cost of ATP via mem- propionate environment. The lower ATPase activity for cells
brane ATPase (Deckers-Hebestreit and Altendorf, 1996; in the stationary phase was attributed to the lower ATP
Kobayashi, 1987; O’Sullivan and Condon, 1999). Thus, an requirement, as cells were not actively growing in the
active ATPase is essential to prevent the acidification of stationary phase. On the other hand, for cells in the
cytoplasm by propionic acid. The effects of propionic acid exponential phase, a lower membrane-bound ATPase activity
concentration on the membrane-ATPase activities of the can be an indication of a more efficient proton pump, as it has
mutant and wild-type strains were studied with cells been reported that the ATPase activity was lower in faster-
harvested at exponential and stationary phases. As shown growing cells than in more slowly growing cells (O’Sullivan
in Figure 9, the activity of ATPase increased with increas- and Condon, 1999).
ing propionic acid concentration in the range between 0 and
10 g/L, but then decreased rapidly to zero at 30 g/L. Membrane Fatty Acid Composition
Surprisingly, the ATPase regained its activity when the
It has been reported that microorganisms such as Sacchar-
propionic acid concentration was 60 g/L and increased with
omyces cerevisiae (Casey and Ingledew, 1986) and Escher-
increasing propionic acid concentration. Also, cells grown in
ichia coli (Ingram, 1976) could increase their tolerance to
the presence of 10 g/L of propionic acid and harvested in the
organic solvents by regulating their membrane lipid compo-
exponential phase had &5% more ATPase activity as
sition in response to environmental stresses. Compared to the
compared with cells grown in the absence of propionic acid
wild type, the membrane fatty acid composition was
(data not shown). These results are consistent with the fact
significantly changed in the mutant, which had more
that the ATPase activity increases as the acid concentration
longer-chain saturated fatty acids (C17:0) and less unsatu-
increases (O’Sullivan and Condon, 1999; Zhu and Yang,
rated fatty acids (C18:1) (Table III), both of which could
2003) or as the extracelluar pH value decreases (Belli and
decrease membrane fluidity and thus might have contributed
Marquis, 1991; Miyagi et al., 1994; O’Sullivan and Condon,
to the increased propionate tolerance.
1999). In all cases studied, the effect of propionic acid on
ATPase was essentially the same for the mutant and wild-type
Morphological Change in Mutant
Cell adaptation and mutation are often accompanied with a
significant change in cell morphology as a response to
environmental perturbations (Jan et al., 2001; Ye et al., 1999).
Propionibacterium acidipropionici has a rod shape; how-
ever, the mutant appeared to be much longer and thinner
than the wild type as observed under SEM (Fig. 10). The
mutant had an average length of 3.2 mm (vs. 1.1 mm for the
wild type) and an average diameter of 0.50 mm (vs. 0.66 mm
for the wild type). Compared to the wild type, the thinner and
longer appearance of the mutant had an &10% increase in its
specific surface area, which should contribute to proportional
increases in substrate uptake and metabolite excretion rates.
This could explain why the mutant might have a more
efficient proton pump and can tolerate a higher propionic acid
concentration. In the FBB high cell density environment with
relatively low glucose and high propionic acid concentra-
tions, the observed morphological change would be a natural
response for cells to adapt and survive. Interestingly, this
dramatic morphological change was permanent and the

Table III. Comparison of membrane fatty acid compositions of


P: acidipropionici wild type and mutant from the FBB.

Fatty Acid Wild type Mutant

C15:0 70.76% 68.5%


C16:0 2.3% 2.9%
Figure 9. Effect of propionic acid on membrane-bound ATPase activity of
C17:0 14.75% 20.06%
P. acidipropionici wild type and mutant from the FBB. (A) Cells in the
C18:1 1.37% 1.24%
exponential phase. (B) Cells in the stationary phase.

334 BIOTECHNOLOGY AND BIOENGINEERING, VOL. 91, NO. 3, AUGUST 5, 2005


propionate tolerant mutants. It was the unique environment in
the FBB that facilitated rapid adaptation of cells and allowed
mutants with higher propionate tolerance to survive when
subjected to adverse conditions. The FBB has also been
successfully used for adapting and screening for acid-tolerant
strains of Clostridium formicoaceticum (Huang et al., 1998)
and Clostridium tyrobutyricum (Zhu and Yang, 2003). The
ability to obtain acid-tolerant mutants in the FBB can be
attributed to: (a) the high cell density (>45 g/L) and viability
(>70%) maintained in the FBB, and (b) distinct physiology
and survivability of immobilized cells resulting from their
direct contact with a solid surface and with each other. None
of these conditions existed in the free-cell fermentation. It
should be noted that the mutant was also tested in free-cell
fed-batch fermentation and gave similar higher propionate
yield and lower succinate yield from glucose, but could not
reach the same high level of propionate concentration as in
the FBB (see Table I). This indicated that besides the
observed mutations in cell morphology, membrane fatty acid
composition, and enzyme activity and sensitivity to propio-
nic acid, there were additional phenotypic changes for cells
immobilized in the FBB that could not be easily duplicated in
the suspension culture.
Immobilization mimics what occurs in nature when cells
adhere and grow on surfaces or within natural structures. The
high cell density in the immobilization system also provides
conditions conducive to cell-to-cell communication. It is
natural for cells to modify their pattern of growth and
Figure 10. Scanning electron micrographs of P. acidipropionici showing replication because of direct contact with a solid surface or
morphological difference between the wild type and the mutant from the with other cells. Immobilization also results in the change of
FBB. (A) Wild-type cells with a short rod shape. (B) Mutant cells with an physicochemical properties of the microenvironment, includ-
elongated slimmer rod shape.
ing the presence of ionic charges, reduced water activity,
altered osmotic pressure, modified surface tension, and cell
mutant preserved this morphology even after numerous confinement. Growth under this situation can induce responses
subculturing in normal growth media for one year. However, at both biochemistry and genetic levels that cause fundamental
this finding is not a surprise as similar observations have also changes in the cells. As a cellular response to the environ-
been reported. Propionibacterium freudenreichii lost its mental stress, immobilization caused changes in cell meta-
original rod shape and became shorter under an extreme bolic pathway and morphology (Krishnan et al., 2001). Cells in
acidic pH of 2.0 (Jan et al., 2001). Lactobacillus plantarum an immobilization system are usually maintained in a non- or
immobilized in chitosan treated polypropylene matrix low-growing but metabolically active state (Paik and Glatz,
experienced a morphological change from a normal rod 1994), which may have allowed them to better survive adverse
shape to a coccoid shape and a metabolic shift from conditions. When subjected to a high ethanol concentration,
homofermentative to heterofermentative (Krishnan et al., immobilized cells were able to maintain a higher cytoplasmic
2001). Lactobacillus rhamnosus immobilized on solid pH or a more efficient proton pump (Groboillot et al., 1994;
support changed from isolated rods to thinner, chain- Loureiro-Dias and Santos, 1990). All these factors might have
linked consortia during continuous lactic acid fermentation contributed to the better adaptation and survivability of cells in
at high dilution rate (Goncalves et al., 1992). a high-density immobilized system such as the FBB.
In the free cell fermentation, both acetate and succinate
production increased with increasing cumulative glucose
Effects of Cell Immobilization in FBB
consumption (see Fig. 3) or propionate concentration.
Clearly, adaptation of P. acidipropionici in the FBB Apparently, free cells grown in suspension would need more
provided an efficient method to obtain a metabolically energy (ATP) when the propionic acid concentration was
robust mutant that can better produce propionic acid higher to maintain their intracellular pH. Thus, there was a
from glucose with a higher concentration and yield. It clear metabolic pathway shift; however, which could not
should be noted that free cell fermentations carried out continue to work at higher concentrations of propionic acid
under similar fed-batch conditions could not achieve the and the fermentation ceased to produce propionic acid
same adaptation effect and was not as effective in obtaining at &52 g/L. On the other hand, for the FBB fermentation,

SUWANNAKHAM AND YANG: ENHANCED PROPIONIC ACID FERMENTATION BY P. ACIDIPROPIONICI MUTANT 335
cells responded differently. Instead of shifting the metabolic Glenner GG. 1977. Formanzans and tetrazolium salts. In: Lillie RD, editor.
pathway, immobilized cells underwent some mutations, H.J. Conn’s biological strains: A handbook on the nature and uses of the
including changes in morphology and membrane fatty acid dyes employed in the biological laboratory. Baltimore, MD: Williams
and Wilkins. p 201–236.
content to increase their tolerance to propionic acid. These Goncalves LMD, Barreto MTO, Xavier AMBR, Carrondo MJT, Klein J.
changes allowed the immobilized cells to tolerate and 1992. Inert supports for lactic acid fermentation—A technical assess-
produce more propioinic acid at a higher concentration. ment. Appl Microbiol Biotechnol 38:305–311.
Goswami V, Srivastava AK. 2000. Fed-batch propionic acid production by
Propionibacterium acidipropionici. Biochem Eng J 4:121–128.
CONCLUSIONS Groboillot A, Boadi DK, Poncelet D, Neufeld RJ. 1994. Immobilization of
cells for application in the food industry. Crit Rev Biotechnol 14:75–107.
Immobilization and adaptation of P. acidipropionici in the Gu Z, Glatz BA, Glatz CE. 1998. Effects of propionic acid on
fibrous-bed bioreactor provided an effective means to obtain propionibacteria fermentation. Enzyme Microb Technol 22:13–18.
a metabolically advantageous mutant with the ability to Gutierrez NA, Maddox IS. 1992. Product inhibition in a nonmotile mutant of
tolerate and produce more propionic acid at higher Clostridium acetobutylicum. Enzyme Microb Technol 14:101–105.
Hsu ST, Yang S-T. 1991. Propionic acid fermentation of lactose by
concentrations and yield from glucose. The mutant had
Propionibacterium acidipropionici: Effects of pH. Biotechnol Bioeng
significant changes in its morphology, membrane lipid 38:571–578.
composition, and key enzymes in the pathways leading to Huang YL, Mann K, Novak JM, Yang S-T. 1998. Acetic acid production
propionic acid and succinic acid. The higher propionate yield from fructose by Clostridium formicoaceticum immobilized in a
was attributed to the higher activity levels of oxaloacetate fibrous-bed bioreactor. Biotechnol Prog 14:800–806.
Huang YL, Wu Z, Zhang L, Cheung CM, Yang S-T. 2002. Production of
transcarboxylase and CoA transferase in the mutant,
carboxylic acids from hydrolyzed corn meal by immobilized cell
while the lower succinate yield in the mutant resulted from fermentation in a fibrous-bed bioreactor. Bioresource Technol 82:51–59.
the lower activity of PEP carboxylase. The increased Huang Y, Yang S-T. 1998. Acetate production from whey lactose using co-
propionate tolerance of the mutant is possibly because of immobilized cells of homolactic and homoacetic bacteria in a fibrous-
changes in its membrane fatty acid composition and its bed bioreactor. Biotechnol Bioeng 60:498–507.
Ingram LO. 1976. Adaptation of membrane lipids to alcohols. J Bacteriol
slimmer morphology making the proton pump more efficient
125:670–678.
as indicated by the lower ATPase activity. The decreased Jan G, Leverrier P, Pichereau V, Boyaval P. 2001. Changes in protein
sensitivity to propionic acid inhibition for the mutant’s synthesis and morphology during acid adaptation of Propionibacterium
oxaloacetate transcarboxylase and CoA transferase also freudenreichii. Appl Environ Microbiol 67:2029–2036.
allowed the cells to survive and continue to produce Jin Z, Yang S-T. 1998. Extractive fermentation for enhanced propionic acid
production from lactose by Propionibacterium acidipropionici. Bio-
propionate even when the propionate concentration was
technol Prog 14:457–465.
high. The higher final product concentration, coupled with Kobayashi H. 1987. Regulation of cytoplasmic pH in streptococci. In: Reizer
the significantly reduced by-product and cell concentrations J, Peterkofsky A, editors. Sugar transport and metabolism in Gram-
in the fermentation broth should facilitate simple separations positive bacteria. London, UK: Ellis Harwood. p 255–269.
for product purification and reduce the overall production Krischke W, Schroder M, Trosch W. 1991. Continuous production of L-lactic
acid from whey permeate by immobilized Lactobacillus casei subsp.
cost of propionic acid by the FBB fermentation process.
casei. Appl Microbiol Biotechnol 34:573–578.
Krishnan Sudha, Gowda LR, Misra MC, Karanth NG. 2001. Physiological
and morphological changes in immobilized L. plantarum NCIM 2084
References cells during repeated batch fermentation for production of lactic
acid. Food Biotechnol 15:193–202.
Allen SHG, Kellermeyer RW, Stjernholm RL, Wood HG. 1964. Purification Lewis VP, Yang S-T. 1992. Continuous propionic acid fermentation by using
and properties of enzymes involved in the propionic acid fermentation. J immobilized Propionibacterium acidipropionici in a novel packed-bed
Bacteriol 87:171–187. bioreactor. Biotechnol Bioeng 40:465–474.
Andersch W, Bahl H, Gottschalk G. 1983. Level of enzymes involved in Loureiro-Dias MC, Santos H. 1990. Effects of ethanol on Saccharomyces
acetate, butyrate, acetone and butanol formation by Clostridium cerevisiae as monitored by in vivo 31P and 13C nuclear magnetic
acetobutylicum. Eur J Appl Microbiol Biotechnol 18:327–332. resonance. Arch Microbiol 153:384–391.
Balamurugan K, Venkata Dasu V, Panda T. 1999. Propionic acid production Maeba P, Sanwal BD. 1969. Phosphoenolpyruvate carboxylase from
by whole cells of Propionibacterium freudenreichii. Bioprocess Eng Salmonella typhimurium, strain LT 2. Methods Enzymol 13:283–288.
20:109–116. McCleary WR, Stock JB, Ninfa AJ. 1993. Is acetyl phosphate a global signal
Belli WA, Marquis RE. 1991. Adaptation of Streptococcus mutans and in Escherichia coli? J Bacteriol 175:2793–2798.
Enterococcus hirae to acid stress in continuous culture. Appl Environ Miyagi A, Ohta H, Kodama T, Fukui K, Kato K, Shimono T. 1994. Metabolic
Microbiol 57:1134–1138. and energetic aspects of the growth response of Streptococcus rattus to
Casey GP, Ingledew WE. 1986. Ethanol tolerance in yeasts. Crit Rev environmental acidification in anaerobic continuous culture.
Microbiol 13:219–280. Microbiology 140:1945–1952.
Deckers-Hebestreit G, Altendorf K. 1996. The F0F1-type ATP synthases of Norton S, Lacroix C, Vuillemard J-C. 1994. Kinetic study of continuous
bacteria: Structure and function of the F0 complex. Annu Rev Microbiol whey permeate fermentation by immobilized Lactobacillus helveticus
50:791–824. for lactic acid production. Enzyme Microb Technol 16:457–466.
Ellman GL. 1959. Tissue sulfhydryl groups. Arch Biochem Biophys 82: O’Sullivan E, Condon S. 1999. Relationship between acid tolerance,
70–77. cytoplasmic pH and ATP and Hþ-ATPase levels in chemostat cultures of
Emde R, Schink B. 1990. Enhanced propionate formation by Propionibac- Lactococcus lactis. Appl Environ Microbiol 65:2287–2293.
terium freudenreichii subsp. freudenreichii in a three-electrode Paik H-D, Glatz BA. 1994. Propionic acid production by immobilized cells
amperometric culture system. Appl Environ Microbiol 56:2771– of a propionate-tolerant strain of Propionibacterium acidipropionici.
2776. Appl Microbiol Biotechnol 42:22–27.

336 BIOTECHNOLOGY AND BIOENGINEERING, VOL. 91, NO. 3, AUGUST 5, 2005


Playne MJ. 1985. Propionic and butyric acids. In: Moo-Young M, editor. Wanner BL, Wilmes-Riesenberg MR. 1992. Involvement of phosphotransa-
Comprehensive Biotechnology, Vol. 3. New York: Pergamon. p 731– cetylase, acetate kinase and acetyl phosphate synthesis in control of the
759. phosphate regulon in Escherichia coli. J Bacteriol 174:2124–2130.
Rickert DA, Glatz CE, Glatz BA. 1998. Improved organic acid production by Wood HG, Jacobson B, Gerwin BI, Northrop DB. 1969. Oxalacetate
calcium alginate-immobilized propionibacteria. Enzyme Microb Tech- transcarboxylase from Propionibacterium. Methods Enzymol 8:215–
nol 22:409–414. 230.
Schulman M, Wood HG. 1975. Succinyl-CoA:propionate CoA-transferase Woskow SA, Glatz BA. 1991. Propionic acid production by a propionic acid-
from Propionibacterium shermanii. Methods Enzymol 35:235–242. tolerant strain of Propionibacterium acidipropionici in batch and semi-
Senthuran A, Senthuran V, Mattiasson B, Kaul R. 1997. Lactic acid continuous fermentation. Appl Environ Microbiol 57:2821–2828.
fermentation in a recycle batch reactor using immobilized Lactobacillus Yang S-T, Huang Y, Hong G. 1995. A novel recycle batch immobilized cell
casei. Biotechnol Bioeng 55:841–853. bioreactor for propionate production from whey lactose. Biotechnol
Silva EM, Yang S-T. 1995. Kinetics and stability of a fibrous-bed bioreactor Bioeng 45:379–386.
for continuous production of lactic acid from unsupplemented acid Yang S-T, Zhu H, Li Y, Hong G. 1994. Continuous propionate production
whey. J Biotechnol 41:59–70. from whey permeate using a novel fibrous bed bioreactor. Biotechnol
Solichien MS, O’Brien D, Hammond EG, Glatz CE. 1995. Membrane-based Bioeng 43:1124–1130.
extractive fermentation to produce propionic and acetic acids: Toxicity Ye K, Shijo M, Miyano K, Shimizu K. 1999. Metabolic pathway of
and mass transfer considerations. Enzyme Microb Technol 17:23–31. Propionibacterium growing with oxygen: enzymes, 13C NMR
Van der Werf MJ, Guettler MV, Jain MK, Zeikus JG. 1997. Environmental analysis, and its application for vitamin B12 production with
and physiological factors affecting the succinate product ratio during periodic fermentation. Biotechnol Prog 15:201–207.
carbohydrate fermentation by Actinobacillus sp. 130Z. Arch Microbiol Zhu Y, Wu Z, Yang S-T. 2002. Butyric acid production from acid hydrolysate
167:332–342. of corn fiber by Clostridium tyrobutyricum in a fibrous-bed bioreactor.
Van Hoek P, Aristidou A, Hahn JJ, Patist A. 2003. Fermentation goes large- Proc Biochem 38:657–666.
scale. CEP 99:37S–42S. Zhu Y, Yang S-T. 2003. Adaptation of Clostridium tyrobutyricum for
Vogel IA. 1978. Vogel’s textbook of quantitative inorganic analysis. London: enhanced tolerance to butyric acid in a fibrous-bed bioreactor.
Longman Group Limited. 756 p. Biotechnol Prog 19:365–372.

SUWANNAKHAM AND YANG: ENHANCED PROPIONIC ACID FERMENTATION BY P. ACIDIPROPIONICI MUTANT 337

Vous aimerez peut-être aussi