(PDF) pH and base counterion affect succinate production in dual-phase Escherichia coli fermentations - DOKUMEN.TIPS (2024)

(PDF) pH and base counterion affect succinate production in dual-phase Escherichia coli fermentations - DOKUMEN.TIPS (1)

ORIGINAL PAPER

pH and base counterion affect succinate production in dual-phaseEscherichia coli fermentations

Shiying Lu Æ Mark A. Eiteman Æ Elliot Altman

Received: 6 March 2009 / Accepted: 11 May 2009 / Published online: 30 May 2009

� Society for Industrial Microbiology 2009

Abstract Succinate production was studied in Esche-

richia coli AFP111, which contains mutations in pyruvate

formate lyase (pfl), lactate dehydrogenase (ldhA) and

the phosphotransferase system glucosephosphotransferase

enzyme II (ptsG). Two-phase fermentations using a defined

medium at several controlled levels of pH were conducted

in which an aerobic cell growth phase was followed by an

anaerobic succinate production phase using 100% (v/v)

CO2. A pH of 6.4 yielded the highest specific succinate

productivity. A metabolic flux analysis at a pH of 6.4 using13C-labeled glucose showed that 61% of the PEP parti-

tioned to oxaloacetate and 39% partitioned to pyruvate,

while 93% of the succinate was formed via the reductive

arm of the TCA cycle. The flux distribution at a pH of 6.8

was also analyzed and was not significantly different

compared to that at a pH of 6.4. Ca(OH)2 was superior to

NaOH or KOH as the base for controlling the pH. By

maintaining the pH at 6.4 using 25% (w/v) Ca(OH)2, the

process achieved an average succinate productivity of

1.42 g/l h with a yield of 0.61 g/g.

Keywords Succinic acid � CO2 � Calcium

Introduction

Succinic acid (succinate) and its derivatives are widely

used as specialty chemicals in foods, pharmaceuticals and

cosmetics [10], and it can serve as a starting material for

many commercially important products [37]. Some anaer-

obic bacteria, such as Anaerobiospirillum succiniciprodu-

cens [13, 14], Actinobacillus succinogenes [11, 31] and

Mannheimia succiniciproducens [15], produce succinate as

the major fermentation product. During anaerobic fer-

mentation, these organisms fix the greenhouse gas CO2 via

carboxylation reactions and convert C3 to C4 metabolites.

Recombinant Escherichia coli can also generate a high

concentration of succinate. For example, an anaerobic

process generates 15.6 g/l succinate with a yield of 0.85 g/g

glucose through overexpression of pyruvate carboxylase in

an alcohol dehydrogenase and lactate dehydrogenase

mutant [25]. Aerobically, E. coli generates 6.7 g/l succi-

nate at a yield of 0.71 g/g glucose, although in this case

CO2 is not fixed [16, 17]). A dual-phase E. coli process

(aerobic growth followed by an anaerobic succinate

production phase) generates nearly 100 g/l succinate at a

productivity of 1.3 g/l h and a yield of 1.1 g/g glucose

[34].

A few studies have investigated the effect of pH on

succinate production. For example, Van der Werf et al.

[32] found that succinate was produced in the pH range of

6.0–7.4 by Actinobacillus sp., but pH was not controlled.

Samuelov et al. [24] reported that for A. succiniciprodu-

cens a pH of 6.2 was better than a pH of 7.2 using NaOH.

Optimal anaerobic succinate production by E. coli was

reported to occur with a 0.2 M sodium phosphate buffer

having a pH of 6.5 at the beginning of the fermentation [1],

though pH was not controlled. Similarly, a 0.2 M sodium

phosphate buffer at an initial pH of 6.5 was found to pro-

vide the optimal initial conditions for succinate production

by Bacteroides fragilis [12]. Other studies select a pH

rather than optimize it, and universally use either Na2CO3

or NaOH for pH control, resulting in sodium ion

accumulation.

S. Lu � M. A. Eiteman (&) � E. Altman

Center for Molecular BioEngineering, University of Georgia,

Athens, GA 30602, USA

e-mail: [emailprotected]

123

J Ind Microbiol Biotechnol (2009) 36:1101–1109

DOI 10.1007/s10295-009-0594-z

(PDF) pH and base counterion affect succinate production in dual-phase Escherichia coli fermentations - DOKUMEN.TIPS (2)

The highest reported succinate concentration has been

generated by E. coli AFP111 and its derivatives using a

dual-phase process [34]. We selected this strain to study the

influence of pH, and also compared its metabolism at two

different values of pH using flux analysis with 13C-labeled

glucose. The formation of succinate requires the use of

base in order to maintain the pH. When NaOH was used in

initial experiments, we observed a decrease in the rate of

succinate formation during the course of the anaerobic

phase. To determine whether the Na? specifically is det-

rimental to succinate formation, we compared three com-

mon bases [NaOH, KOH and Ca(OH)2] and elucidated

their effects as base counterions.

Materials and methods

Bacterial strain

Escherichia coli AFP111 (F? k- rpoS396 (Am) rph-1

DpflAB::Cam ldhA::Kan ptsG) was used in this study

[3, 5].

Media and fermentation conditions

All fermentations used a defined medium containing

(per liter): 40.0 g glucose, 3.0 g citric acid, 3.0 g Na2

HPO4�7H2O, 8.00 g KH2PO4, 8.00 g (NH4)2HPO4, 0.20 g

NH4Cl, 0.75 g (NH4)2SO4, 0.84 g NaHCO3, 1.00 g

MgSO4�7H2O, 10.0 mg CaCl2�2H2O, 0.5 mg ZnSO4�7H2O, 0.25 mg CuCl2�2H2O, 2.5 mg MnSO4�H2O, 1.75 mg

CoCl2 6H2O, 0.12 mg H3BO3, 1.77 mg Al2(SO4)3�xH2O,

0.5 mg Na2MoO4�2H2O, 16.1 mg Fe(III) citrate, 20 mg

thiamine HCl and 2 mg biotin.

Dual-phase fermentations operating in batch mode with

an initial volume of 1.2 l in 2.5-l fermentors (Bioflow III,

New Brunswick Scientific, Edison, NJ) were inoculated

using 50 ml grown for 10–12 h to an optical density

(OD600) of about 2–4 in the same medium in 250-ml shake

flasks. Oxygen-enriched air as necessary was sparged at

1.0 l/min with an agitation of 200–500 rpm to maintain the

dissolved oxygen (DO) above 40% as measured by an on-

line probe (Mettler-Toledo Process Analytical Instruments,

Wilmington, MA). During growth, the pH was controlled

at 7.0 with 20% (w/v) NaOH, and the temperature was

maintained at 37�C. When the OD600 reached about 20, the

aerobic growth phase was terminated by switching the inlet

gas composition to 100% (v/v) CO2. Simultaneously, the

total flow rate was reduced to 500 ml/min (dry basis, 0�C

and 1 atm), the agitation reduced to 200 rpm, and 120 ml

of 550 g/l glucose was added. During this anaerobic phase,

the pH was controlled with either 25% (w/v) NaOH, 25%

(w/v) KOH or 25% (w/v) Ca(OH)2 as indicated.

Analysis of glucose and products

Samples were diluted by 1% H2SO4 as necessary and

centrifuged (10,0009g for 10 min at 4�C) and the super-

natant analyzed for glucose, succinate, pyruvate, acetate

and ethanol by high performance liquid chromatography

(HPLC) as previously described [6]. When calcium succi-

nate was generated, the total succinate (dissolved plus

precipitate) was determined by diluting the sample with 1%

H2SO4 and sonicating prior to analysis.

For metabolic flux analysis involving [1-13C] glucose,

dual-phase fermentations were repeated as described

above except the initial volume was 0.6 l in 1.0-L fer-

mentors (Bioflow III, New Brunswick Scientific, Edison,

NJ), the inoculum volume was reduced to 25 ml, and the

initial glucose concentration was 30 g/l. The glucose

concentration was monitored using a glucose analyzer

(YSI 2700 SELECTTM, Yellow Springs Instrument, Inc,

Yellow Springs, OH). When the glucose concentration

reached about 2 g/l, the system was switched to anaerobic

conditions by sparging CO2. When the glucose concen-

tration reached about 1 g/l, 10 ml of 450 g/l [1-13C]

glucose (99%, Cambridge Isotope Laboratories, Andover,

MA) was added into the fermentor, and the pH was

reduced within a couple minutes to the desired value with

20% (v/v) H2SO4. Samples were then collected every

30 min, centrifuged at 0�C (10,0009g for 10 min) and the

supernatants stored at -4�C for later LC-MS and NMR

analyses.

LC-MS analysis of succinate

The supernatant was filtered by a 0.2-lM syringe filter

(Cole-Parmer Instrument Co., Vernon Hills, IL) prior to

LC-MS analysis. The liquid chromatography was per-

formed at 25�C (140B Solvent Delivery System, Applied

Biosystems, Foster City, CA) using a C18 column (Kro-

masil, 250 9 1 mm, 5-lm particles, 100-A pore space,

Keystone Scientific, Inc., Bellefonte, PA). The mobile

phase at 50 ll/min was a gradient of aqueous acetic acid

(0.1% w/v, pH 3.23) and methanol. Electrospray ionization

mass spectrometry was performed using a single quadru-

pole instrument with an electrospray ion source (PE Sciex

API 1 Plus, Concord, ON, Canada). The operating condi-

tions were: nebulizer gas (N2) at a flow rate of 0.6 l/min

and pressure of 30 psi, curtain gas (N2) at a flow rate of

0.8 l/min and interface temperature 50�C. Data were

acquired in negative ion mode with a capillary voltage of

3,500 V. Mass peak heights were determined using the

BioTool Box Version B software package (Applied Bio-

systems/PE Sciex, Foster City, CA). The relative concen-

trations of M ? 0, M ? 1, M ? 2 and M ? 3 (M ? 0

represents succinate without 13C label having a m/z of 117,

1102 J Ind Microbiol Biotechnol (2009) 36:1101–1109

123

(PDF) pH and base counterion affect succinate production in dual-phase Escherichia coli fermentations - DOKUMEN.TIPS (3)

etc.) were calculated by the fractions of each peak heights

after correcting for naturally occurring 13C (1.109%) and18O (0.20%) [22].

Metabolic modeling and flux analysis

An overall flux balance was developed using stoichiome-

tric analysis and the pseudo-steady-state condition for

intracellular metabolites [27]. The balance equations for

E. coli AFP111 metabolism have been developed previously

[34], and this model was further extended by including the

pentose phosphate pathway (Fig. 1). Flux partitions (/) were

defined as fractional fluxes (m) into one of two key branches.

The fraction of flux into the pentose phosphate pathway (/PPP)

is defined as

/PPP ¼v3

v2 þ v3

The fraction of flux through anaplerotic carboxylation at

the PEP node (/PPC) is defined as:

/PPC ¼v13

v13 þ v14

The model contained 27 fluxes and 21 metabolites (Fig. 1

and Appendix A). Since insignificant growth occurs during

the anaerobic process [34], glucose consumption and

product formation rates were essentially constant (e.g.,

Fig. 2); thus, the process exhibited a metabolic pseudo-

steady state. Although five fluxes are known, the balance

equations represent an underdetermined set. In order to

calculate the optimal solution for the flux model, LC-MS

results with 1-13C-glucose were used. Specifically, we

wrote analogous balance equations for isotopomers of each

metabolite [28]. We calculated the distribution of succinate

isotopomers for a given set of fluxes and calculated the

mole fraction (Xcalc) of each isotopomer i. These calculated

values were compared with the mole fractions of each

isotopomer observed from the LC-MS results (Xobs). The

optimal solution for the flux model was then determined by

minimizing the weighted sum of squared residuals over the

four mass peaks [33]:

error ¼

ffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffiffi

X

3

i¼0

Xcalc;i � Xobs;i

� �2

s2i

v

u

u

t

where s is the standard deviation of the observed mole

fraction determined by three LC-MS analyses of each

sample. The isotopomer analysis cannot determine whether

flux occurred through the malic enzyme and/or pyruvate

oxidase because these pathways do not alter the13C-labeled distribution compared, respectively, with PEP

carboxylase and acetate formation via acetyl CoA. There-

fore, the model did not include these two pathways.

NMR analysis of succinate

The sample supernatant was filtered by a 0.2-lM syringe

filter, and then the filtered supernatant was mixed with 15%

volume of deuterium dioxide (D2O) for NMR analysis.

Proton-coupled 13C NMR spectra (500 NMR spectroscopy;

Varian Inc., Palo Alto, CA) at 125.7 MHz were obtained with

the following spectral parameters: 45� pulses, 31.4-kHz

spectral width and 45-s relaxation delay. Field stabilization

was achieved by locking on the D2O frequency. 13C chemical

shift assignments for succinate were determined by compa-

rison with the natural abundance standard.

NMR was not used to calculate fluxes, but was

employed to complement the LC-MS results. Specifically,

since the LC-MS results did not provide information about

the position of the label, we were able to use NMR to

observe the total 13C enrichment at C-1 position (S1) and

that at C-2 position of succinate (S2). The enrichment ratio

of S2/S1 was calculated from the ratio of the peak area of

the methylene group to the carboxyl group of succinate.

This observed enrichment ratio was compared to the

enrichment ratio calculated from the optimized metabolic

flux model. We also calculated the redox ratio as the fluxes

through NAD(P)H formation steps in the pathways divided

by fluxes through NAD(P)H consumption steps.

We chose for all metabolic flux analyses the interval

between the time 1-13C-glucose was added and the time

that the succinate concentration reached about 4 g/l

(approximately 2 h).

Results

Effect of pH on succinate production

E. coli AFP111 accumulates significant amounts of succi-

nate during an anaerobic non-growth phase after growing to

a high cell density under aerobic conditions at a pH of 7.0.

During succinate generation, CO2 is incorporated into the

central metabolism through the action of the enzyme PEP

carboxylase (PPC) [8]. We first examined the effect of pH

on succinate formation. We controlled the pH at a constant

level in the range of 5.8–7.0 using 25% Ca(OH)2 as the

neutralizing agent. Figure 2 shows the generation of several

products during the anaerobic process when the pH was

maintained at 6.4. Over the course of 14–16 h, succinate

was generated to 25–30 g/l, pyruvate to 7–10 g/l, ethanol to

about 1 g/l and acetate to less than 1 g/l. For each pH

studied, the glucose consumption rate (qG), the volumetric

succinate productivity (Qs), specific succinate productivity

(qS) and mass product yields of succinate, acetate, ethanol

and pyruvate were calculated during the 14 h of the

anaerobic phase, and the values reported are the means of

J Ind Microbiol Biotechnol (2009) 36:1101–1109 1103

123

(PDF) pH and base counterion affect succinate production in dual-phase Escherichia coli fermentations - DOKUMEN.TIPS (4)

two to three experiments (Table 1). In the pH range of 5.8–

6.4, qG and qS increased with increasing pH, but the yields

of succinate and other products did not significantly change.

When controlled pH was greater than 6.6, qG and qS

decreased. Fermentations in which the pH was controlled at

6.4 resulted in the highest specific succinate productivity.

Moreover, the volumetric succinate productivity at a pH of

6.4 or 6.6 remained high throughout the course of the

anaerobic production phase (about 1.2–1.8 g/l h), resulting

in the highest mean succinate productivity (Table 1). For a

pH above 6.6, the productivity declined over the course of

the anaerobic phase (from about 2.0 to 0.5 g/l h). Since pure

(acidic) CO2 was sparged into the fermenter during the

anaerobic phase and three acid products were formed, base

was required to maintain the pH (Fig. 3). Indeed, an unac-

ceptably large quantity of base was needed above a pH of

about 6.6. The observed reduction in specific productivity,

which would account for any ‘‘dilution effect,’’ demon-

strates that the cells generated a maximal rate of succinate at

the intermediate pH. Because of these results, a pH con-

trolled at 6.4 during the anaerobic production phase was

selected for the subsequent study.

Metabolic flux analysis

Succinate is formed through two pathways: the reductive

arm of the tricarboxylic acid cycle (TCA) via the anaple-

rotic enzyme PPC and the glyoxylate shunt [35]. To

glucose-6P

glucose

fructose-6P

glyceraldehyde-3P

ribulose-5P

sedoheptulose-7P

erythrose-4P

ribose-5P

xylulose-5P

glyceraldehyde-3P

PEP

pyruvate

isocitrate

succinate

fumarate

malate

glyoxylate

oxaloacetate

CO2

4

6

52

3

acetate

citrate

10

11

9

8

7

12

16

13

CO 2

15

17 18

3P-glycerate

fructose-1,6P2

acetyl CoA

CO2

19

succinate (ext)

acetate (ext)

ethanol (ext)

20

14

21

26

2325

27

24

22

pyruvate (ext)

1

Fig. 1 Biochemical pathways

for the synthesis of succinate

from glucose in E. coli. Not all

enzymatic steps or

intermediates are shown. Key

enzymes in the pathways are as

follows: (1) gluco*kinase,

(2) phosphoglucoisomerase;

(3) 6-phosphogluconate

dehydrogenase; (4)

phosphopentose epimerase; (5)

phosphopentose epimerase; (6)

transketolase; (7) transaldolase;

(8) transketolase; (9)

phosphofructokinase; (10)

fructose biphosphate aldolase;

(11) glyceraldehyde 3-

phosphate dehydrogenase and

phosphoglycerate kinase; (12)

phosphoglycerate mutase and

enolase; (13) PEP carboxylase;

(14) pyruvate kinase; (16)

pyruvate dehydrogenase

complex; (17)

phosphoacetyltransferase; (18)

acetate kinase; (19)

acetaldehyde dehydrogenase

and alcohol dehydrogenase;

(20) citrate synthase; (21)

aconitase; (22) isocitrate lyase;

(23) malate synthase; (24)

malate dehydrogenase; (25)

fumarase; (26) fumarate

reductase

1104 J Ind Microbiol Biotechnol (2009) 36:1101–1109

123

(PDF) pH and base counterion affect succinate production in dual-phase Escherichia coli fermentations - DOKUMEN.TIPS (5)

understand whether pH affects the distribution of these two

pathways, we compared the metabolic fluxes at a pH of 6.4

to that at a pH of 6.8. We first determined the flux distri-

bution at a pH of 6.4. Table 3 compares the observed mass

distribution of succinate and the calculated mass distribu-

tion by the optimal metabolic model (i.e., least error as

defined in ‘‘Materials and methods’’). Two methods were

used to validate the metabolic model. First, we compared

the enrichment ratio S2/S1 observed in the NMR results

with that ratio predicted from the metabolic model obtained

from independent LC-MS results (Table 2). In addition, we

calculated the redox balance based on the metabolic model,

and the value was very close to 1 (Table 2). The resulting

metabolic fluxes for the process operating at a pH of 6.4

showed that 93% of the 1.25 mol succinate formed per

mole glucose was generated via the reductive arm of the

TCA cycle and 7% via the glyoxylate shunt (Fig. 4a). At

the PEP node, /PPC was 0.61 (i.e., about 61% of PEP

partitioned to the reductive branch of the TCA cycle).

Although 39% of the PEP formed pyruvate, most of the

pyruvate ultimately became by-products (external pyru-

vate, acetate and ethanol), and only 12% of the pyruvate

became succinate via the glyoxylate shunt (either directly

or from malate).

During succinate accumulation CO2 was released via

6-phosphogluconate dehydrogenase and pyruvate dehy-

drogenase (or pyruvate oxidase) but sequestered via PPC

(Fig. 1). At a pH of 6.4, the net CO2 consumption rate was

1.10 mmol/g h. The overall stoichiometric coefficient for

CO2 (i.e., the ratio of net CO2 consumption to glucose

consumption) was 0.62 (Fig. 4a).

We similarly determined the metabolic fluxes at a pH of

6.8 (Table 2). Although the higher pH significantly reduced

the glucose consumption rate and succinate production rate

(Fig. 4b), it did not alter the carbon partitioning compared

to a pH of 6.4: succinate yield was 1.24 mol/mol glucose,

/PPP of 0.15 and /PPC of 0.62 (Table 2). Also, 38% of

the intermediate PEP formed pyruvate, and about 10% of

the pyruvate became succinate via the glyoxylate shunt; the

CO2 stoichiometric coefficient was 0.59 (Fig. 4b).

Time (h)

0 2 4 6 8 10 12 14 16 18

Glu

cose

(g/

L)

10

20

30

40

50

60

70

Pro

duct

s (g

/L)

5

10

15

20

25

30

Fig. 2 Production of succinate (filled square), pyruvate (opendiamond), ethanol (open triangle) and acetate (open square) from

glucose (filled circle) during the anaerobic non-growth production

phase for E. coli AFP111 using 25% Ca(OH)2 to control pH at 6.4

Table 1 The volumetric productivity (Q), specific consumption or production rates (q) and product mass yields (Y) of E. coli AFP111 during

14 h of an anaerobic non-growth phase using Ca(OH)2 as the base to control pH

pH qG (mg/g h) QS (g/l h) qS (mg/g h) YS (g/g) YA (g/g) YP (g/g) YE (g/g)

5.8 108.7a 0.67a 76.5a 0.71a 0.01a 0.24a 0.00a

6.0 189.3b 1.02b 126.0b 0.67a 0.01a 0.21a 0.02a

6.2 192.6b 1.18bc 125.7b 0.65a 0.02a 0.24a 0.02a

6.4 284.6c 1.42c 174.3c 0.61a -0.01a 0.24a 0.03a

6.6 257.4bc 1.49c 169.3c 0.66a 0.00a 0.16a 0.03a

6.8 166.3b 0.99b 119.4ab 0.72a 0.01a 0.19a 0.03a

7.0 85.7d 0.36d 47.4d 0.55b 0.06b 0.05b 0.02a

Different letters were statistically significantly different at the 90% confidence level

G glucose, S succinate, A acetate, P pyruvate, E ethanol

pH

5.6 5.8 6.0 6.2 6.4 6.6 6.8 7.0 7.2

Bas

e C

onsu

mpt

ion

(mL)

200

400

600

800

1000

1200

1400

Fig. 3 Volume of 25% Ca(OH)2 consumption during 14 h of an

anaerobic non-growth production phase for E. coli AFP111 under

different levels of controlled pH

J Ind Microbiol Biotechnol (2009) 36:1101–1109 1105

123

(PDF) pH and base counterion affect succinate production in dual-phase Escherichia coli fermentations - DOKUMEN.TIPS (6)

Effect of base counterion

Using a pH of 6.4 and 100% CO2 in the gas phase, we next

examined the effect of the type of neutralizing agent used

on succinate production. Three different bases were com-

pared: 25% KOH, 25% NaOH or 25% Ca(OH)2. The

effects of K? and Na? appeared to be similar, with the

succinate productivity declining during the anaerobic phase

(from about 1.7 to 0.3 g/l h). The difference in volume

used (Table 3) was due to the unequal molar base con-

centrations, and the total moles of KOH or NaOH added

during the 14-h anaerobic phase were similar. When

Ca(OH)2 was used for pH control, some calcium succinate

precipitated, and the volumetric succinate productivity

remained high during 14 h of an anaerobic production

phase (about 1.2–1.8 g/l h), resulting in the highest mean

Table 2 Comparison of the mass distributions of succinate formed by E. coli AFP111 as observed by mass spectrometry and as calculated by the

optimal metabolic model

pH uPPP uPPC M ? 0 M ? 1 M ? 2 M ? 3 (S2/S1)a R/O

6.4 0.12 0.61 Observed 0.585 ± 0.006 0.399 ± 0.009 0.014 ± 0.008 0.002 ± 0.002 19.56

Calculated 0.591 0.384 0.022 0.003 18.32 0.97

6.8 0.15 0.62 Observed 0.580 ± 0.008 0.395 ± 0.009 0.018 ± 0.005 0.009 ± 0.006 18.76

Calculated 0.576 0.398 0.022 0.003 18.79 1.06

The enrichment ratio (S2/S1) was observed from NMR results. The redox ratio R/O was calculated from the optimal metabolic model

Observed values are shown as mean ± standard deviation from three analysesa The standard deviation for NMR measurements was 4–10%

glucose-6P

glucose

fructose-6P

glyceraldehyde-3P

ribulose-5P

sedoheptulose-7P

erythrose-4P

ribose-5P

xylulose-5P

glyceraldehyde-3P

PEP

pyruvate

isocitrate

succinate

fumarate

malate

glyoxylate

oxaloacetate

CO2

acetate

citrate

CO2

3P-glycerate

fructose-1,6P2

acetyl CoA

CO2

succinate (ext)

acetate (ext)

ethanol (ext)

pyruvate (ext)

1.77

1.56

0.21

0.140.07

0.070.070.07

1.70

3.40

3.47

3.47

1.34

2.13 0.52

0.31

0.82

0.31

0.350.08

0.08

0.08

0.080

2.05

2.13

2.13

2.21

glucose-6P

glucose

fructose-6P

glyceraldehyde-3P

ribulose-5P

sedoheptulose-7P

erythrose-4P

ribose-5P

xylulose-5P

glyceraldehyde-3P

PEP

pyruvate

isocitrate

succinate

fumarate

malate

glyoxylate

oxaloacetate

CO2

acetate

citrate

CO2

3P-glycerate

fructose-1,6P2

acetyl CoA

CO2

succinate (ext)

acetate (ext)

ethanol (ext)

pyruvate (ext)

0.83

0.71

0.12

0.080.04

0.040.040.04

0.79

1.58

1.62

1.62

0.62

1.00 0.23

0.26

0.39

0.26

0.070.03

0.03

0.03

0.03

0.97

1.00

1.00

1.03

A B

Fig. 4 Metabolic fluxes (mmol/g h) during an anaerobic non-growth production phase of E. coli AFP111 at a pH of 6.4 (a) and a pH of 6.8 (b)

1106 J Ind Microbiol Biotechnol (2009) 36:1101–1109

123

(PDF) pH and base counterion affect succinate production in dual-phase Escherichia coli fermentations - DOKUMEN.TIPS (7)

succinate productivity (qS) (Table 3). The small reduction

in succinate productivity over the course of the anaerobic

phase for Ca(OH)2 can be attributed to a dilution of fer-

menter contents (i.e., adding base to a non-growing popu-

lation). Twenty-five percent Ca(OH)2 and 25% NaOH have

similar OH- molar concentrations (0.67 M for Ca(OH)2

and 0.63 M for NaOH). However, during the 14-h anaer-

obic phase about twice the amount of Ca(OH)2 was con-

sumed compared to NaOH (Table 3), which is consistent

with the higher qS for Ca(OH)2. The values of qG and YS

using Ca(OH)2 were also higher than those at NaOH or

KOH as a base counterion (Table 3).

Discussion

Between the pH range of 5.8–6.8, the specific succinate

formation and glucose consumption rates achieved their

maximum values, although the fermentation product yields

were not significantly affected by the external pH

(Table 1). Most aerobic and facultatively anaerobic bac-

teria stringently regulate the cytoplasmic pH (pHi), and the

pHi of E. coli is unaffected by large variations in the

medium pH (pHex) [2]. Olsen et al. [21] found that the pHi

of E. coli (7.0–8.0) did not vary significantly over the pHex

ranging from 5.5 to 8.0. However, the pH gradient

(DpH = pHi - pHex) is approximately 1.5 at a pHex of 5.5

and decreases with increasing pHex, ultimately reaching 0

at a pHex of 8.0 [21]. A large DpH is associated with more

active transport of H? and other ions across the membrane

to maintain pH homeostasis [2]. One might anticipate that

the increased maintenance requirement resulting from

reduced pH might reduce the product yield. Although no

product yield reduction was observed, the burden of

maintaining DpH at low pH may be responsible for the

reduced rates of substrate utilization and succinate

formation.

The succinate formation rate did not increase indefi-

nitely with increasing pH; it decreased above a pH of 6.4.

The quantity of buffer required to maintain the pH

increased with increasing pH (Fig. 3). Thus, when the pH

was above 6.4, the ionic strength increased more quickly

during the succinate formation phase, an effect that may

have been detrimental to the rate of succinate formation.

Two mechanisms may exist that result in the observed

optimum pH: a DpH effect that reduces succinate forma-

tion at low pH and an ionic strength effect that reduces the

succinate formation rate at higher pH. Interestingly, the

optimal pH for succinate productivity observed in this

study (6.4) corresponds closely with the pKa of the car-

bonic acid/bicarbonate equilibrium (6.35). Our optimal pH

for succinate production by E. coli AFP111 near 6.4 is also

consistent with other succinate-producing bacteria,

including A. succiniciproducens at a pH of 6.2 [24] and

Bacteroides fragilis at a pH of 6.5 [12].

The results demonstrate that 13C-labeling can be suc-

cessfully applied to calculate metabolic fluxes during non-

growth succinate production. For this process, the reductive

branch of the TCA cycle was the most important pathway

for succinate formation. This result is consistent with other

succinate-producing E. coli strains, including a ptsG

mutant TUQ2 [36] and E. coli strains that overexpress

Lactococcus lactis pyruvate carboxylase [26]. Surprisingly,

the flux distribution/carbon distribution was not signifi-

cantly different between a pH of 6.4 and 6.8. In particular,

the pH of the medium, which in this pH range would have a

large impact on the CO2

HCO�3 ratio [7, 29], did not affect

the flux partition at the PEP node (Fig. 4). Similar to our

results with E. coli, the distribution of fermentation prod-

ucts by Actinobacillus sp. 130z [32] did not differ in the pH

range of 6.0–7.4. The key enzyme PPC uses bicarbonate as

the form of CO2 [20], and flux through this pathway could

be expected to depend on the external availability of

CO2

HCO�3 . Indeed, in the presence of limiting concen-

trations of CO2, E. coli AFP111 formed less succinate (data

not shown); however, in this study 100% CO2 was used.

Under these circ*mstances, the yield and flux results sup-

port the conclusion that CO2 was not limiting despite the

dissociation between the dissolved gas and bicarbonate in

this pH range. Consequently, under these non-limiting

conditions, increasing the total quantity of CO2 (i.e., CO2

plus HCO�3 ) as a result of increasing the pH would not lead

to an increase in the PPC flux and associated succinate

formation. Furthermore, as the previous research described

above would suggest, pHi is likely unchanged over the pH

range studied, and pH would therefore not affect the

Table 3 The volumetric productivity (Q), specific consumption or production rates (q) and product mass yields (Y) of E. coli AFP111 during

14 h of an anaerobic non-growth phase using three different bases to control pH at 6.4

Base (25% w/v) qG (mg/g h) QS (g/l h) qS (mg/g h) YS (g/g) YP (g/g) YA (g/g) YE (g/g) Vbase (ml)

Ca(OH)2 248.6 ± 31.8 1.42 ± 0.19 174.3 ± 40.2 0.61 ± 0.13 0.24 ± 0.04 -0.01 ± 0.02 0.03 ± 0.02 135 ± 18

KOH 177.9 ± 19.0 0.88 ± 0.07 95.3 ± 5.8 0.54 ± 0.08 0.37 ± 0.01 -0.01 ± 0.02 0.01 ± 0.03 113 ± 7

NaOH 176.2 ± 28.2 0.99 ± 0.19 101.6 ± 23.9 0.57 ± 0.04 0.21 ± 0.08 0.02 ± 0.01 0.00 ± 0.02 66 ± 15

Data were presented as mean ± standard deviation from two to three replications

G glucose, S succinate, A acetate, P pyruvate, E ethanol

J Ind Microbiol Biotechnol (2009) 36:1101–1109 1107

123

(PDF) pH and base counterion affect succinate production in dual-phase Escherichia coli fermentations - DOKUMEN.TIPS (8)

activities of the various enzymes in the glucose to succinate

metabolic pathways.

The results demonstrate that Ca2? is superior to Na? or

K? as a base counterion to control the pH during the fer-

mentation. Although the specific cause for this difference is

unknown, the calcium succinate has a solubility of only

11.8 g/l at 40�C [19], which is far lower than sodium

succinate or potassium succinate. Thus, the use of Ca(OH)2

would have the double benefit of removing both the cation

Ca2? and the anion succinate from the solution. The

observed removal of succinate by the precipitation of cal-

cium succinate could relieve the inhibition of PPC [4] and

isocitrate lyase [18], two key enzymes in succinate for-

mation [35]. Removal of the cation would reduce osmotic

stress. At the end of the processes studied, the concentra-

tion of Na? or K? was about 0.5 M, while the remaining

concentration of dissociated Ca2? would have been only

about 0.18 M. High ion strength leads to osmotic stress,

and under aerobic conditions E. coli exports some Na? as a

response [23]. However, under anaerobic conditions, Na?

extrusion activity may decrease [30]. Moreover, in our

study the high ionic strength only occurred under anaerobic

non-growth conditions, potentially limiting the cellular

ability to respond to that stress. A recent study on lactate

generation similarly demonstrated that monovalent cations

reduce acid formation by E. coli compared to Ca2? [38].

Gouesbet et al. [9] found that about 0.24 M KCl (osmo-

larity of 820 mOsm) had inhibitory and repressive effects

on anaerobic enzymes and the corresponding genes in

E. coli due to osmotic stress.

In summary, although the pH does affect the production

rate, the pH within the range of 5.8–6.8 does not affect

yield or the flux distribution of products in a two-phase

succinate production process using E. coli. The negative

impact of osmotic stress during succinate accumulation can

at least partly be alleviated by using calcium as the product

counterion.

Acknowledgments Financial support from the US Department of

Energy (DE-FG26-04NT42126) is gratefully acknowledged. We

acknowledge Drs. G.P. Wylie, T.M. Andacht and D.R. Phillips for

NMR and MS analyses. We also acknowledge S.A. Lee for technical

assistance and Y. Zhu for helpful discussion.

Appendix A: Metabolic reactions for succinate

production

m1 Glucose ? ATP � glucose 6-phosphate ? ADP

m2 Glucose 6-phosphate � fructose 6-phosphate

m3 Glucose 6-phosphate ? 2 NADP? ? H2O � ribulose

5-phosphate ? CO2 ? 2 NADPH ? 2 H?

Appendix continued

m4 Ribulose 5-phosphate � ribose-5-phosphate

m5 Ribulose 5-phosphate � xylulose 5-phosphate

m6 Ribose 5-phosphate ? xylulose 5-phosphate � glyceraldehyde

3-phosphate ? sedoheptulose 7-phosphate

m7 Glyceraldehyde 3-phosphate ? sedoheptulose-7-phosphate �

fructose 6-phosphate ? erythrose 4-phosphate

m8 Xylulose 5-phosphate ? erythrose 4-phosphate � fructose

6-phosphate ? glyceraldehydes 3-phosphate

m9 Fructose 6-phosphate ? ATP � fructose 1,6-

bisphosphate ? ADP

m10 Fructose 1,6-bisphosphate � dihydroxyacetone

phosphate ? glyceraldehyde 3-phosphate

m11 Glyceraldehyde 3-phosphate ? NAD? ? ADP ? Pi �

3-phosphoglycerate ? NADH ? H? ? ATP

m12 3-Phosphoglycerate � PEP ? H2O

m13 PEP ? CO2 ? H2O � oxaloacetate ? Pi

m14 Phosphoenolpyruvate ? ADP � pyruvate ? ATP

m15 Pyruvate (intracellular) � pyruvate (extracellular)

m16 Pyruvate ? NAD? ? CoA � acetyl-

CoA ? CO2 ? NADH ? H?

m17 Acetyl-CoA ? ADP ? Pi � acetate ? CoA ? ATP

m18 Acetate (intracellular) � acetate (extracellular)

m19 Acetyl-CoA ? 2 NADH ? 2 H?� ethanol ? 2 NAD?

m20 Oxaloacetate ? acetyl-CoA ? H2O � citrate ? CoA

m21 Citrate � isocitrate

m22 Isocitrate � glyoxylate ? succinate

m23 Glyoxylate ? acetyl-CoA ? H2O � malate ? CoA

m24 Oxaloacetate ? NADH ? H?� malate ? NAD?

m25 Malate � fumarate ? H2O

m26 Fumarate ? NADH ? H?� succinate ? NAD?

m27 Succinate (intracellular) � succinate (extracellular)

References

1. Agarwal L, Isar J, Meghwanshi GK, Saxena RK (2006) A cost

effective fermentative production of succinic acid from cane

molasses and corn steep liquor by Escherichia coli. J Appl

Microbiol 100:1348–1354. doi:10.1111/j.1365-2672.2006.02894.x

2. Booth IR (1985) Regulation of cytoplasmic pH in bacteria.

Microbiol Rev 49:359–378

3. Chatterjee R, Millard CS, Champion K, Clark DP, Donnelly MI

(2001) Mutation of the ptsG gene results in increased production

of succinate in fermentation of glucose by Escherichia coli. Appl

Environ Microbiol 67:148–154. doi:10.1128/AEM.67.1.148-154.

2001

4. Corwin LM, Fanning GR (1968) Studies of parameters affecting

the allosteric nature of phosphoenolpyruvate carboxylase of

Escherichia coli. J Biol Chem 243:3517–3525

5. Donnelly MI, Millard CS, Clark DP, Chen MJ, Rathke JW (1998)

A novel fermentation pathway in an Escherichia coli mutant

producing succinic acid, acetic acid, and ethanol. Appl Biochem

Biotechnol 70–72:187–198. doi:10.1007/BF02920135

6. Eiteman MA, Chastain MJ (1997) Optimization of the ion-

exchange analysis of organic acids from fermentation. Anal Chim

Acta 338:69–75. doi:10.1016/S0003-2670(96)00426-6

1108 J Ind Microbiol Biotechnol (2009) 36:1101–1109

123

(PDF) pH and base counterion affect succinate production in dual-phase Escherichia coli fermentations - DOKUMEN.TIPS (9)

7. Frahm B, Blank HC, Cornand P, Oelssner W, Guth U, Lane P,

Munack A, Johannsen K, Portner R (2002) Determination of

dissolved CO2 concentration and CO2 production rate of mam-

malian cell suspension culture based on off-gas measurement.

J Biotechnol 99:133–148. doi:10.1016/S0168-1656(02)00180-3

8. Gokarn RR, Eiteman MA, Altman E (2000) Metabolic analysis of

Escherichia coli in the presence and absence of carboxylating

enzymes phosphoenolpyruvate carboxylase and pyruvate car-

boxylase. Appl Environ Microbiol 66:1844–1850. doi:10.1128/

AEM.66.5.1844-1850.2000

9. Gouesbet G, Abaibou H, Wu LF, Mandrand-Berthelot MA,

Blanco C (1993) Osmotic repression of anaerobic metabolic

systems in Escherichia coli. J Bacteriol 175:214–221

10. Guettler MV, Jain MK, Soni BK (1998) Process for making

succinic acid, microorganisms for use in the process and methods

of obtaining the microorganisms. US patent 5,723,322, 3 March

1998

11. Guettler MV, Rumler D, Jain MK (1999) Actinobacillus succin-ogenes sp. nov., a novel succinic-acid-producing strain from the

bovine rumen. Int J Syst Bacteriol 49:207–216

12. Isar J, Agarwal L, Saran S, Saxena RK (2006) Succinic acid

production from Bacteroides fragilis: process optimization and

scale up in a bioreactor. Anaerobe 12:231–237. doi:10.1016/

j.anaerobe.2006.07.001

13. Lee PC, Lee WG, Kwon S, Lee SY, Chang HN (2000) Batch and

continuous cultivation of Anaerobiospirillum succiniciproducensfor the production of succinic acid from whey. Appl Microbiol

Biotechnol 54:23–27. doi:10.1007/s002530000331

14. Lee PC, Lee WG, Lee SY, Chang HN (2001) Succinic acid

production with reduced by-product formation in the fermenta-

tion of Anaerobiospirillum succiniciproducens using glycerol as a

carbon source. Biotechnol Bioeng 72:41–48. doi:10.1002/1097-

0290(20010105)72:1\41::AID-BIT6[3.0.CO;2-N

15. Lee SJ, Song H, Lee SY (2006) Genome-based metabolic engi-

neering of Mannheimia succiniciproducens for succinic acid

production. Appl Environ Microbiol 72:1939–1948. doi:10.1128/

AEM.72.3.1939-1948.2006

16. Lin H, Bennett GN, San KY (2005) Genetic reconstruction of the

aerobic central metabolism in Escherichia coli for the absolute

aerobic production of succinate. Biotechnol Bioeng 89:148–156.

doi:10.1002/bit.20298

17. Lin H, Bennett GN, San KY (2005) Metabolic engineering of

aerobic succinate production systems in Escherichia coli to

improve process productivity and achieve the maximum theo-

retical succinate yield. Metab Eng 7:116–127. doi:10.1016/

j.ymben.2004.10.003

18. MacKintosh C, Nimmo HG (1988) Purification and regulatory

properties of isocitrate lyase from Escherichia coli ML308.

Biochem J 250:25–31

19. Miczynski ZN (1886) Uber die Bestimmung der Loslichkeit

einiger Sauren und Salze der Oxalsaurereihe in Wasser bei ver-

schiedenen Temperaturen. Monatsh Chem 7:255–272. doi:

10.1007/BF01516575

20. O’Leary MH (1982) Phosphoenolpyruvate carboxylase: an

enzymologist’s view. Annu Rev Plant Physiol 33:297–315. doi:

10.1146/annurev.pp.33.060182.001501

21. Olsen K, Budde B, Siegumfeldt H, Rechinger K, Jakobsen M,

Ingmer H (2002) Noninvasive measurement of bacterial intracel-

lular pH on a single-cell level with green fluorescent protein and

fluorescence ratio imaging microscopy. Appl Environ Microbiol

68:4145–4147. doi:10.1128/AEM.68.8.4145-4147.2002

22. Rosman KJR, Taylor PDP (1998) Isotopic compositions of

the elements. Pure Appl Chem 70:217–230. doi:10.1351/pac

199870010217

23. Sakuma T, Yamada N, Saito H, Kakegawa T, Kobayashi H

(1998) pH dependence of the sodium ion extrusion systems in

Escherichia coli. Biochim Biophys Acta 1363:231–237. doi:

10.1016/S0005-2728(97)00102-3

24. Samuelov N, Lamed R, Lowe S, Zeikus JG (1991) Influence of

CO2–HCO3- levels and pH on growth, succinate production, and

enzyme activities of Anaerobiospirillum succiniciproducens.

Appl Environ Microbiol 57:3013–3019

25. Sanchez AM, Bennett GN, San KY (2005) Efficient succinic acid

production from glucose through overexpression of pyruvate

carboxylase in an Escherichia coli alcohol dehydrogenase and

lactate dehydrogenase mutant. Biotechnol Prog 21:358–365. doi:

10.1021/bp049676e

26. Sanchez AM, Bennett GN, San KY (2006) Batch culture char-

acterization and metabolic flux analysis of succinate-producing

Escherichia coli strains. Metab Eng 8:209–226. doi:10.1016/

j.ymben.2005.11.004

27. Savinell JM, Palsson BO (1992) Network analysis of intermedi-

ary metabolism using linear optimization. I. Development of

mathematical formalism. J Theor Biol 154:421–454. doi:10.1016/

S0022-5193(05)80161-4

28. Schmidt K, Carlsen M, Nielsen J, Villadsen J (1997) Modeling

isotopomer distributions in biochemical networks using isotopo-

mer mapping matrices. Biotechnol Bioeng 55:831–840. doi:

10.1002/(SICI)1097-0290(19970920)55:6\831::AID-BIT2[3.0.

CO;2-H

29. Stumm W, Morgan J (1996) Dissolved carbon dioxide. In:

Aquatic chemistry, John Wiley, New York, p 192

30. Trchounian A, Kobayashi H (1999) Fermenting Escherichia coliis able to grow in media of high osmolarity, but is sensitive to the

presence of sodium ion. Curr Microbiol 39:109–114. doi:

10.1007/s002849900429

31. Urbance S, Pometto AIII, Dispirito A, Denli Y (2004) Evaluation

of succinic acid continuous and repeat-batch biofilm fermentation

by Actinobacillus succinogenes using plastic composite support

bioreactors. Appl Microbiol Biotechnol 65:664–670. doi:

10.1007/s00253-004-1634-2

32. Van der Werf MJ, Guettler MV, Jain MK, Zeikus JG (1997)

Environmental and physiological factors affecting the succinate

product ratio during carbohydrate fermentation by Actinobacillussp. 130z. Arch Microbiol 167:332–342. doi:10.1007/s0020300

50452

33. Van Dien SJ, Strovas T, Lidstroml ME (2003) Quantification of

central metabolic fluxes in the facultative methylotroph methy-lobacterium extorquens AM1 using 13C-label tracing and mass

spectrometry. Biotechnol Bioeng 84:45–55. doi:10.1002/bit.

10745

34. Vemuri GN, Eiteman MA, Altman E (2002) Succinate production

in dual-phase Escherichia coli fermentations depends on the time

of transition from aerobic to anaerobic conditions. J Ind Micro-

biol Biotechnol 28:325–332. doi:10.1038/sj.jim.7000250

35. Vemuri GN, Eiteman MA, Altman E (2002) Effects of growth

mode and pyruvate carboxylase on succinic acid production by

metabolically engineered strains of Escherichia coli. Appl

Environ Microbiol 68:1715–1727. doi:10.1128/AEM.68.4.1715-

1727.2002

36. Wang Q, Chen X, Yang Y, Zhao X (2006) Genome-scale in silico

aided metabolic analysis and flux comparisons of Escherichiacoli to improve succinate production. Appl Microbiol Biotechnol

73:887–894. doi:10.1007/s00253-006-0535-y

37. Zeikus JG, Jain MK, Elankovan P (1999) Biotechnology of

succinic acid production and markets for derived industrial

products. Appl Microbiol Biotechnol 51:545–552. doi:10.1007/

s002530051431

38. Zhu Y, Eiteman MA, DeWitt K, Altman E (2007) hom*olactate

fermentation by metabolically engineered Escherichia colistrains. Appl Environ Microbiol 73:456–464. doi:10.1128/AEM.

02022-06

J Ind Microbiol Biotechnol (2009) 36:1101–1109 1109

123

(PDF) pH and base counterion affect succinate production in dual-phase Escherichia coli fermentations - DOKUMEN.TIPS (2024)
Top Articles
Latest Posts
Article information

Author: Carmelo Roob

Last Updated:

Views: 6270

Rating: 4.4 / 5 (65 voted)

Reviews: 80% of readers found this page helpful

Author information

Name: Carmelo Roob

Birthday: 1995-01-09

Address: Apt. 915 481 Sipes Cliff, New Gonzalobury, CO 80176

Phone: +6773780339780

Job: Sales Executive

Hobby: Gaming, Jogging, Rugby, Video gaming, Handball, Ice skating, Web surfing

Introduction: My name is Carmelo Roob, I am a modern, handsome, delightful, comfortable, attractive, vast, good person who loves writing and wants to share my knowledge and understanding with you.