Appl. Environ. Microbiol. - Applied and Environmental Microbiology

AEM Accepts, published online ahead of print on 18 July 2014
Appl. Environ. Microbiol. doi:10.1128/AEM.01391-14
Copyright © 2014, American Society for Microbiology. All Rights Reserved.
1
Sorbic acid and acetic acid have distinct effects on the electrophysiology and metabolism of
2
Bacillus subtilis
3
4
5
J.W.A. van Beilena, M.J. Teixeira de Mattosb, K.J. Hellingwerfb, S. Brula#
6
7
Molecular Biology & Microbial Food Safety, Swammerdam Institute for Life Sciences,
8
University of Amsterdam, The Netherlandsa.
9
Department of Molecular Microbial Physiology, Swammerdam Institute for Life Sciences,
10
University of Amsterdam, The Netherlandsb.
11
12
Running Head: Physiology of weak acid stress in Bacillus subtilis
13
14
#
corresponding author: S. Brul, [email protected]
15
16
17
1
18
Abstract
19
Sorbic and acetic acid are amongst the most commonly used weak organic acid preservatives
20
to improve the microbiological stability of foods. Both have a similar pKa value but sorbic
21
acid is a far more potent preservative. Weak organic acids are most effective at low pH.
22
Under these circumstances, they are assumed to diffuse across the membrane as neutral
23
undissociated acids. We show here that the level of initial intracellular acidification depends
24
on the concentration of undissociated acid and less on the nature of the acid. Recovery of the
25
internal pH depends on the presence of an energy source, but acidification of the cytosol
26
causes a decrease in glucose flux. Furthermore, sorbic acid is a more potent uncoupler of the
27
membrane potential than acetic acid. Together these effects may also slow down the rate of
28
ATP synthesis significantly and may thus (partially) explain sorbic acid’s effectiveness.
2
29
Introduction
30
Various small weak organic acids (WOA) have been used as food preservatives for a very
31
long period of time. These weak acids slow down growth of various spoilage bacteria, yeasts
32
and moulds without overt undesired effects on taste or being toxic to the consumer. The
33
undissociated states of the WOA preservatives are more effective in slowing growth than the
34
dissociated form, although the latter may have some level of toxicity. As such, WOAs are
35
most effective when applied at low pH values, below their pKa value (1, 2). Under these
36
conditions, the neutral acid is assumed to diffuse across the plasma membrane and dissociate
37
in the cytosol which generally has a higher pH. In this way, the proton gradient over the
38
membrane is depleted and the anion may accumulate to potentially toxic levels inside the
39
cell. This is known as the classical “weak-acid preservative” theory (3). Commonly used
40
WOA preservatives include sorbic and acetic acid, which have a similar pKa value of 4.76 but
41
a dissimilar octanol: water partition coefficient (log Kow) of 1.33 and -0.17 respectively (4).
42
This means that at a particular pH and the same total concentration, concentrations of both
43
undissociated acids are the same, but sorbic acid has a higher affinity for a hydrophobic
44
(membrane) environment. Sorbic acid is clearly the more potent preservative of the two, but
45
the exact reason why is still not fully clear (5, 6).
46
It is important to distinguish the different modes of action that WOAs may have on cell
47
physiology. The classical “weak-acid preservative” theory (3) only assumes entry of the
48
undissociated acid, dissociation in the cytosol and cytosolic acidification. While this is
49
sometimes described as uncoupling, we will use this latter phrase for compounds that shuttle
50
protons across the membrane and are thus protonophoric uncouplers. Others have also
51
pointed to potential toxicity of accumulated anions (1, 7, 8). If we assume that only the
52
undissociated acid diffuses across the membrane, it follows from the Henderson–Hasselbalch
53
equation (9) that ∆
, and thus the internal concentration of the anion may
3
54
become very high when a high ΔpH remains present. Also, specific binding of sorbate to
55
cysteine (10) has been shown and is proposed as an explanation for the higher toxicity of
56
sorbic acid.
57
Several resistance mechanisms against WOAs have been reported for yeasts like
58
Saccharomyces cerevisiae, Zygosaccharomyces baillii and Z. rouxii. These organisms induce
59
the expression of H+-ATPases to regulate their cytosolic pH. S. cerevisiae uses a dedicated
60
ATP binding cassette (ABC) transporter (Pdr12) to prevent accumulation of the anion (6).
61
Also S. cerevisiae plasma membrane components are likely to play an important role in the
62
modulation of the influx of lipophilic weak organic acids (11, 12). Furthermore, Z. rouxii and
63
Aspergillus niger have been shown to degrade sorbic acid to 1,3-pentadiene (13, 14).
64
The responses and potential resistance mechanisms of bacteria against weak acids have not
65
been so well-described as for yeasts (15, 16). The level of growth reduction has been
66
modelled some 30 years ago (1) and the effect of sorbic acid on the membrane potential in
67
Escherichia coli membrane vesicles has been described (17). More recently Ter Beek et al.
68
(2) performed a microarray study on B. subtilis exposed to sorbic acid. Their results showed a
69
broad transcriptomic response resembling a pattern typical for cells responding to nutrient
70
limitation. The authors observed an upregulation of genes encoding potential efflux pumps as
71
well as genes involved in remodelling of the plasma membrane.
72
Bacillus subtilis has been the Gram positive model organism for decades, because it is
73
Generally Recognized As Safe (GRAS), it is genetic accessibility and has a fully sequenced
74
genome. It also forms heat resistant spores and is as such a recognized spoilage organism
75
(18). Spores of several related bacterial species are of great concern to the food industry
76
because they are highly resistant to most preservation techniques and, once germinated, can
77
cause food spoilage through growth of vegetative cells that may produce toxins (16).
4
78
To comprehensively elucidate the physiological effects of WOAs on B. subtilis, we measured
79
the effects of sorbic and acetic acid on the chemical (ΔpH) and electrical (ΔΨ) component of
80
the proton motive force (PMF). With a depleted PMF, we speculated that the cell might alter
81
its energy needs in terms of glucose consumption and availability of its terminal electron
82
acceptor (i.e. O2), which were hence determined. We assessed the rate and extent of change
83
in pHi caused by these two weak acids using B. subtilis cells that were either directly exposed
84
to both WOAs or had been pre-exposed to sorbic acid and re-exposed to WOAs. The latter
85
experiment was done because we inferred from the previously collected micro-array data that
86
cells elicit an adaptive response to these compounds (2).
87
88
Methods
89
General growth conditions and strains
90
For general purposes B. subtilis PB2 strains were grown in Lysogeny Broth (LB). For weak
91
acid stress experiments and fluorescence measurements, B. subtilis strains were grown in
92
defined liquid medium (M3G; (19) set at pH 5.5, 6.4 (buffered with 80 mM MES), as well as
93
7.0 and 7.4 (buffered with 80 mM MOPS). The medium contained 5 mM glucose, 10 mM
94
glutamate, and 10 mM NH4Cl as carbon and nitrogen sources. All cultures were grown at
95
37°C, under continuous agitation at 200 rpm. The wild-type strain used (B. subtilis PB2) was
96
obtained from C.W. Price and A. Ter Beek (2). The strain expressing IpHluorin (B. subtilis
97
PB2 PptsG-IpHluorin) was constructed as described (20). When required, 50 µg/ml
98
spectinomycin was added to the medium.
99
100
Calibration of pHluorin and Internal pH measurements
101
The internal pH was measured as described (20). All strains were grown in M3G at pH = 6.4.
102
For rapid-exposure WOA stress experiments, potassium sorbate (K-S) and potassium acetate
5
103
(K-Ac) were used at 250 mM, dissolved in M3G medium without glucose. Of these solutions,
104
2-10 µl were injected into the cell suspensions at 310 µl/s using the injector of the FluoStar
105
Optima (BMG Labtech, Germany). As a control experiment, the same concentrations of
106
either KCl or NaCl were injected into the microtiter wells. For 50% and 80% growth
107
inhibition experiments, 3 and 11 mM K-sorbate or 25 and 80 mM K-acetate were used,
108
respectively.
109
110
Cell counts and protein measurements
111
In order to compare results per cell from the different experiments, cell counts were
112
performed with a CASY counter (Roche, Germany) equipped with a 60 µm tube. The number
113
of cells counted were 4-10x104 cells per ml. Protein concentrations were determined using a
114
BCA kit (Thermo Scientific) according to the manufacturer’s instructions. Finally, the OD600
115
of the cultures was measured with a FluoStar Optima.
116
117
Membrane potential measurements
118
B. subtilis PB2 and B. subtilis PptsG-IpHluorin were grown as described in M3G, pH = 6.4, to
119
OD600 = 0.4. Cells were harvested by centrifugation and resuspended in 1/10 volume M3G
120
without glucose. To inhibit growth and protein synthesis between experiments, 10 µg/ml
121
chloramphenicol was added. Cells were stored at 37°C. The membrane potential (ΔΨ) was
122
measured using a Tetra Phenyl Phosphonium ion (TPP+) electrode (World Precision
123
Instruments, Inc., USA) filled with 1 mM TPP+. All measurements were performed in a
124
warmed (37°C), 2-ml measuring cell, containing a TPP+ electrode, reference electrode and an
125
oxygen sensor (see below). The cell suspension was stirred magnetically and aerated by a
126
continuous flow of compressed air, so that at least 120 µM oxygen was present when Δψ was
127
measured. One ml of cell suspension was added to 1 ml medium with glucose; subsequently,
6
128
weak organic acids and a TPP+ calibration series (1, 1, 2, 4 and 8 µl of 1 mM TPP+) were
129
added. The addition of weak organic acid did affect the offset, but not the slope of the
130
calibration response. Stress conditions tested were: 3 and 11 mM K-S and 25 and 80 mM K-
131
Ac. These conditions were compared to non-exposed cells. The membrane potential was
132
calculated as described (21–24) using de-energized cells (incubated for 10 min at 70°C) as
133
reference for non-specific binding of the probe.
134
The electric potential was calculated using the following equation (21, 24):
∆
135
1
·
1
1
Where
2.303 ·
136
137
Δψ = transmembrane electrical potential (mV, ψin – ψout)
138
Z = Conversion factor
139
F = Faraday’s constant (96.6 J mV-1 mol-1)
140
R = universal gas constant (8.31 J K-1 mol-1)
141
n = charge of the translocated ion
142
T = absolute temperature (K)
143
C0 = Probe concentration in the medium without addition of cells
144
Ce = Extracellular probe concentration
145
fcm = Ratio of fractional cytoplasmic membrane and intracellular volume
146
Kcm = Cytoplasmic membrane partition coefficient
147
x = Fractional internal volume
148
7
149
The factor fcmKcm was determined by a probe binding assay (21) to be 14. In this (simplified)
150
approach one assumes that the amount of TPP+ bound to extra-cytoplasmic components of
151
the cell (like parts of the cell wall) is equal to the amount of probe that binds to intracellular
152
components like nucleic acids. At the end of the experiment, 1.5-ml samples were collected
153
for protein quantification as described above. Three biological replicates were measured for
154
each condition.
155
156
Oxygen consumption measurements
157
Oxygen consumption rates were measured simultaneously with the Δψ, in the same
158
measuring cell using a Neofox fiber optic oxygen sensor (Ocean Optics Inc., USA). Oxygen
159
concentrations were measured every second from the maximally aerated to the fully oxygen
160
depleted state. The slope of the straight part of the plot (at least 15 seconds) was used to
161
derive the oxygen consumption rate. At the end of the experiment, 1.5-ml samples were
162
collected for protein quantification as described above. Rates were normalized to protein
163
content.
164
165
Glucose consumption and metabolite measurements
166
B. subtilis PB2 and B. subtilis PptsG-IpHluorin were grown as described in M3G, pH = 6.4, to
167
OD600 = 0.8. The cultures were split and exposed to different stress conditions (3 and 11 mM
168
K-sorbate or 25 and 80 mM K-Ac), and 10 mM glucose was added to each culture. A high
169
OD600 was required to observe a significant decrease in glucose concentrations within the
170
timeframe of the experiment.
171
Samples, taken every 30 minutes, were snap frozen in liquid nitrogen. The protein
172
concentration of each sample was measured as described above. Samples were further
173
processed for HPLC analysis; a 1 ml sample was mixed with 100 µl 35% perchloric acid and
8
174
subsequently 55 µl 7 M KOH was added. Filtered supernatants were analysed for glucose
175
consumption levels by the cells as well as the presence of fermentation end-products.
176
Glucose, succinate, lactate, acetate, 2,3-butanediol and ethanol contents were determined by
177
HPLC (LKB) with a REZEX organic acid analysis column (Phenomenex) at 45°C with 7.2
178
mM H2SO4 as the eluent, using an RI 1530 refractive index detector (Jasco) and AZUR
179
chromatography software for data integration. All measurements were performed with two
180
biological replicates. From the obtained metabolite fluxes, a carbon balance was calculated
181
with the assumption that similar molar amounts of CO2 were produced per amount of O2
182
consumed plus similar molar amounts of CO2 were produced per mole of acetate produced
183
and per mole of 2,3-butanediol, 2 moles of CO2 were produced. Oxygen consumption was
184
corrected for directly related acetate and 2,3-butanediol production.
%
100% · 2
5
, ,
/6
185
186
Results
187
188
Internal pH during growth under WOA stress remains low
189
Weak organic acids have been shown to lower the internal pH of microorganisms (25, 26).
190
Thus, we measured pHi during growth with a number of WOA stresses. To reduce growth
191
rate by approximately 50%, 3 mM K-sorbate or 25 mM K-acetate was used. To reduce
192
growth rate by 85%, 10 and 80 mM were used, respectively.
193
Growth and pHi were monitored for 6 hours (Figures 1A and 1B respectively). The internal
194
pH of non-stressed cultures dropped from pHi = 7.5 to 7.2 during this time. The internal pH
195
of sorbic acid-stressed cultures continued to drop from t = 0 until the end of the experiment.
196
With 80 mM K-Ac, this is not seen, and pHi remained stable around pH = 7.3. These results
9
197
show that there is considerable acidification of the cell during growth and that there is no
198
recovery of the pHi during extended exposure to WOAs, not even when growth resumes.
199
200
Weak organic acids cause rapid drop in internal pH
201
Long term (6 h) exposure to WOAs as described above, showed no long-term recovery of the
202
pHi of the cells. The largest drop in pHi upon weak acid exposure occurs within the first
203
minutes after addition. To investigate the influx rate of weak organic acid preservatives and
204
their dependence on the metabolic activity of the cell, we investigated the short term effects
205
of WOA injections into the medium.
206
Before acid injection, the pHi of starved wild-type B. subtilis was 7.30 ± 0.05. Injection of
207
either KCl or NaCl at identical concentrations as the weak acids had no effect on pHi within
208
one minute after injection (not shown). Upon glucose injection, the pHi rapidly increased to
209
7.5 within 1 minute (not shown). When WOAs were injected, the pH dropped to its lowest
210
point within 1-4 s. In starved cells, this pH remained stable, but in the presence of glucose,
211
pHi recovered quickly to a new equilibrium (Figures 2A-D). The acidification of starved cells
212
was fitted with a first order kinetic equation
0
1
·
213
where (Bi) is the amplitude factor (ΔpH/mM), indicating the intracellular buffering capacity,
214
[HA] is the concentration of undissociated weak acid (in mM), (ki) is the rate constant (s-1)
215
and pHi(0) is the offset (pHi at t=0, before injection). Using this equation, values for (Bi) and
216
(ki) were determined for sorbic and acetic acid. The value for ΔpHi or the amplitude of
217
acidification has a linear relation with the concentration of undissociated acid in the medium
218
(Figure 4). These data (with starved cells) show that sorbic acid causes a similar change in
219
pHi as acetic acid does. The rate of acidification is high and similar for both acids, 1.29 ±
10
220
0.09 s-1 for sorbic acid and 1.27 ± 0.33 s-1 for acetic acid. However, with a measuring
221
frequency of one per second, this rate is most likely set by our detection system and the actual
222
rate is likely higher. Cells pre-exposed to sorbic acid had similar values as non-stressed cells
223
for Bi and ki for either WOA, showing no sign of adaptation at this level (Supplementary
224
table 1 and our unpublished data).
225
With glucose added to the cultures, recovery of the pHi started immediately after the injection
226
of weak acid. This curve too was fitted, but with an additional factor describing recovery.
0
1
1
227
Where Br is the amplitude of recovery and kr is the rate constant of the pH recovery.
228
Recovery kinetics do not seem to change between unexposed and pre-exposed cultures. There
229
is however a clear difference between sorbic and acetic acid. The pHi of acetic acid stressed
230
cultures recovers to a higher pH than sorbic acid stressed cultures (Figure 3 and
231
supplementary table 1). The amplitude of recovery also appears to have a linear relation with
232
the acid concentration. The acquired constants allow predictions for both acidification and
233
subsequent recovery (Supplementary figure 1).
234
235
Sorbic acid affects both ΔpH and ΔΨ
236
Proton translocation by the electron transport chain through the cell membrane results in both
237
a gradient in the chemical potential of protons (i.e. ΔpH), as well as an electrical potential
238
(ΔΨ). This electrochemical proton gradient exerts an inward-directed proton motive force or
239
PMF (27, 28). The PMF can drive protons back into the cell via the F1F0-ATPase for ATP
240
synthesis, and is also required for various membrane transport processes and for flagellar
241
rotation. The PMF is defined as (e.g. (29):
∆
242
·
∆ψ
·∆
Where
11
2.303 ·
243
Δp= Proton motive force, PMF, (mV)
244
ΔµH+ = transmembrane electrochemical proton potential (J/mol)
245
F = Faraday’s constant (96.6 J mV-1 mol-1)
246
Δψ = transmembrane electrical potential (mV, ψin – ψout)
247
Z = Conversion factor
248
ΔpH = transmembrane pH gradient (pHin – pHout)
249
R = universal gas constant (8.31 J K-1 mol-1)
250
T = absolute temperature (K)
251
n = charge of the translocated ion
252
253
In our experiments, which were performed at 37°C, Z = 61.4 mV. Our measurements with
254
TPP+ show that under non-stressed conditions with extracellular pH = 6.4, Δψ = 103 ± 9 mV
255
and the pHi = 7.78 ± 0.05. This thus gives a Z*ΔpH value of 84 ± 3 mV and a total PMF of
256
188 ± 9 mV (Figure 5). This value for Δψ is similar to earlier reported values (23, 30) and the
257
pHi value is slightly higher (30). Sorbic acid has a severe effect on Δψ, reducing it by 64%
258
even at 3 mM K-S. Acetic acid, however, does not deplete Δψ as strongly. With 25 mM K-
259
Ac, the Δψ is 17% lower and with 80 mM only a 45% reduction is seen. It appears that sorbic
260
acid acts more as an uncoupler than acetic acid does.
261
Together, Δψ - Z*ΔpH comprise the proton motive force. In total, WOA stress causes a
262
dissipation from 188 ± 9 mV in non-stressed cells to 83 ± 5 mV with 11 mM K-S and 108 ± 7
263
mV with 80 mM K-Ac. This too shows that sorbic acid has a stronger effect on the
264
electrochemical gradient for protons than acetic acid does at concentrations that lead to
265
similar growth inhibition (Supplementary table 2).
266
12
267
Oxygen consumption is reduced upon addition of high WOA concentrations
268
Because of the depleted PMF, the cell may experience energy depletion stress, since the F1F0-
269
ATPase depends on the proton gradient for its activity. It is possible that the cell tries to
270
compensate for the decrease in PMF by increasing the flux through the electron transfer chain
271
and increasing proton pumping activity. This also requires a terminal electron acceptor (e.g.
272
O2). We therefore measured the oxygen levels during our PMF experiments and made sure
273
that at least 120 µM O2 was measured when Δψ was measured. Non-stressed cells consumed
274
oxygen at a rate of 7.4 mmol s-1 (mg protein)-1. Sorbic acid stress reduced the rate of oxygen
275
consumption to 3.9 and 1.5 mmol s-1 (mg protein)-1 with 3 and 11 mM of K-sorbate
276
respectively. 25 mM of K-acetate had a non-significant effect on the oxygen consumption
277
rate, but 80 mM reduced it to 1.1 mmol s-1 (mg protein)-1 (Figure 6). These effects took place
278
immediately after addition of the acid. These observations show that the cell consumes less
279
O2 when faced with weak organic acid stress.
280
281
Glucose metabolism is affected by weak organic acid stress
282
Because weak organic acids partially dissipate the PMF, we speculated that B. subtilis might
283
alter its metabolism
284
Glucose consumption was highest for non-stressed cells, reaching 7.8 mmol s-1 (mg protein)-
285
1
. Addition of weak acids lowered glucose consumption rates, with 80 mM K-Ac resulting in
286
1.45 mmol s-1 (mg protein)-1 (Figure 7 and supplementary table 3).
287
Acetate was produced under all conditions tested and already present in low amounts at the
288
start of the experiment. Non-stressed cultures produced acetate at the highest rate, together
289
with the K-sorbate exposed cells. Addition of K-sorbate lowered the synthesis rate of acetate.
290
Acetic acid-stressed cells had the lowest rate of acetate production. These cultures re-directed
13
291
fermentation routes by switching to 2,3-butanediol fermentation. Some 2,3-butanediol was
292
also found in the final sample of non-stressed cells, but not with sorbic acid-stressed cells.
293
294
Carbon flux
295
When confronted with a lack of electron acceptor, B. subtilis employs a mixed acid
296
fermentation, and has been shown to produce lactate, acetate, acetoin, ethanol, 2,3-butanediol
297
and succinate as fermentation products when grown on glucose (31), but also produces a lot
298
of acetate as a result of overflow metabolism (32). In our experiments, only acetate and 2,3-
299
butanediol were found.
300
Although batch-cultures are not ideally suited for accurate carbon flux determinations, our
301
data do seem to be in line with reported values for wild-type B. subtilis (33). With the
302
assumptions made previously regarding O2 utilization, our data (Supplementary table 3) for
303
WOA stressed cells show reasonably close C-balances for all stresses, apart from 25 mM K-
304
Ac, which appears to use about 3.5x as much O2 as estimated based on carbon fluxes based
305
solely on glycolysis and fermentation.
306
The calculated qATP based on qAcetate and q2,3-Butanediol (qATP = 2*qAcetate + 2*q2,3-
307
Butanediol) through glycolysis and fermentation displayed a similar WOA concentration
308
dependent behaviour as the qGlucose. That is, increasing concentrations of weak acid cause a
309
reduction in fluxes. This behaviour seems analogous to the effect on pHi, which depends
310
more on the concentration than the nature of the acid (as shown above). Indeed, there appears
311
to be a linear relation between pHi and qGlucose (Figure 8), as well as between pHi, and the
312
calculated amounts of ATP generated via production of qAcetate and q2,3-Butanediol (our
313
unpublished calculations). When looking at correlations between Δψ and qGlucose, such a
314
correlation does not exist for both WOA. Only for sorbic acid might a linear correlation be
315
observed, but more experiments will be needed to confirm this relationship. The relation
14
316
between qO2 and decrease in pHi seems rather constant, while a decreasing qO2 seems to
317
correlate with a decreasing ΔΨ. However, more experiments will be needed for an accurate
318
description.
319
320
321
Discussion and conclusions
322
To reduce the growth rate to a similar level, more acetic acid is required than sorbic acid at an
323
identical external pH value. Both sorbic and acetic acid lower pHi and reduce the growth rate
324
of B. subtilis. When growing with WOA stress, the internal pH is lower than that of cells not
325
exposed to WOAs during the early exponential phase. Even though in the first minute after
326
WOA exposure some recovery takes place, the internal pH does not seem to recover fully of
327
this acidification, even after 6 h of growth. Given that the pKa for both weak acids is almost
328
equal, this makes sorbic acid a more effective preservative.
329
Homeostasis of the pHi is crucial for cell physiology. Optimal proton extrusion depends
330
clearly on the presence of an energy source (i.e. glucose) in the medium, be it that recovery of
331
the pHi does not reach the original values. It appears that a new ΔpH equilibrium is
332
established within a few minutes, i.e. before most transcriptomic changes take effect. The
333
level of cytosolic acidification after weak acid injection into the medium is very similar for
334
both acids without glucose present. This suggests that in starved cells, an equilibrium for acid
335
dissociation is established. Since the only factor measured in this regard is pH or [H+], the
336
anion accumulation level is unknown. It can, however, be estimated if it is assumed that
337
∆
338
reduce anion concentrations, but high enough to allow metabolism. As is clear from the data,
339
the ability to restore pHi differs between sorbic acid and acetic acid-treated cells.
. The cell has to establish a new equilibrium with a ΔpH low enough to
15
340
For sorbic acid, the inability to recover pHi during acid stress reflects the ability of sorbate to
341
act as a classic uncoupler, shuttling protons over the membrane whereas the less lipophilic
342
acetate is believed to do this to a lesser extent. The latter is corroborated by the greater effect
343
that sorbic acid has on the membrane potential, while acetic acid only carries bulk volume
344
protons across the membrane until a steady state is reached, allowing Δψ to remain relatively
345
unaffected. Even though some authors have pointed to the similar effects of the uncoupler
346
2,4-dinitrophenol and sorbic acid on growth and metabolism (10, 34, 35), to our knowledge
347
no quantitative measurements of the effect that sorbic acid has on Δψ have been reported thus
348
far. While relatively little research has been published on the effects that WOA preservatives
349
have on the membrane potential of microorganisms (7, 17, 30, 36, 37), the depletion of Δψ
350
may have a plethora of effects. The consequences may range from reduced transporter
351
activity (54) to destabilization of the bacterial cytoskeleton (55), although we have not seen
352
evidence for the latter effect (our unpublished results). Anion efflux from the cytosol would
353
be driven by the remaining Δψ. This effect was shown clearly for the permeant ion picrate,
354
which mainly acts as an uncoupler in everted cell membranes when a high Δψ was
355
established (38). It has also been observed that lipophilic compounds such as ethanol can
356
stimulate leakage of protons over the membrane of S. cerevisiae (39, 40). However, recent
357
results show this not to be the case for sorbic or acetic acid (6).
358
Other commonly used preservatives such as lactic, benzoic and propionic acid are also
359
expected to diffuse and dissociate in agreement with the Henderson – Hasselbalch equation
360
(9). Future experiments should establish to what extent these and other preservatives have
361
uncoupling properties, and what the quantitative effect of Δψ depletion is on growth rate. It is
362
likely that their ability to cross the membrane barrier (log KOW) into the aqueous phase will
363
play a role, as seems to be the case with sorbic and acetic acid (41, 42).
16
364
With a low pHo, the balance of the PMF is shifted towards ΔpH, so the effectiveness of weak
365
organic acids at low pH is two-fold: The concentration of undissociated acid is higher and
366
they act on the dominant component of the PMF. In the search for food preservatives that are
367
also active near neutral pH, the food industry should consider preferably uncoupler-like
368
molecules similar to or better than sorbate. At an even higher pH, permeable cations might be
369
useful to deplete the membrane potential. In food with low glucose concentrations, the
370
recovery of pHi might be less, hence less growth is observed with lower preservative
371
concentrations.
372
We observed a decrease in O2 consumption rate with WOA-stressed cells. This suggests that
373
there are fewer electrons to feed the electron transport chain and that metabolism may be
374
affected by acidification of the cytosol. Because weak organic acids partially dissipate the
375
PMF, the amount of ATP that the cells can generate through the F1F0-ATPase is considerably
376
smaller. So, to produce a sufficient amount of ATP to restore pH homeostasis and proliferate,
377
the cell might alter its glucose metabolism to generate ATP through substrate level
378
phosphorylation. Because increased energy needs have been observed in WOA stressed yeast
379
(43, 44) and a starvation-like response was reported for sorbic acid-stressed B. subtilis (2), we
380
measured the glucose consumption rate as well as the production of fermentation products.
381
Glucose metabolism slows down in the presence of WOA, depending on the concentration of
382
undissociated acid, as has also been described for E. coli and Saccharomyces cerevisiae (25,
383
45, 46). This results in a linear relation between pHi and qGlucose. The observation that
384
acetic acid continues to be produced when cells are exposed to sorbic acid (2 mmol h-1 (mg
385
protein)-1 with 11 mM K-sorbate), may be an extra stress factor for cells growing under these
386
conditions, and may be part of the explanation for the observed continued decrease in pHi
387
during sorbic acid stress. Addition of K-sorbate lowered the synthesis rate of acetate, possibly
388
due to growth reduction or a decrease in glycolytic activity
17
389
Glycolytic flux can be controlled, amongst other factors by pH (47), affecting
390
phosphofructokinase (Pfk) activity (48), or specifically in the case of Bacillus species,
391
phosphoglycerate mutase activity (49), which is also very sensitive to pH changes. Glucose is
392
generally assumed to be the preferred carbon and energy source for B. subtilis which is taken
393
up via the phosphoenolpyruvate-sugar phosphotransferase system (PTS). It is subsequently
394
converted into fructose 1,6-bisphosphate (FBP) by Pfk. FBP is required for phosphorylation
395
of CcpA, one of the main repressors in carbon catabolite repression (CCR) in B. subtilis. This
396
way, FBP availability causes CCR (50). Also, CCR inhibits expression of citric acid cycle-
397
enzymes when glucose concentrations are high. When a decrease of the internal pH reduces
398
the activity of Pfk, the concentration of FBP is reduced, CcpA (a global regulator (inhibitor)
399
of many carbon metabolism routes(51)) activity is reduced and CCR is released. This may
400
explain how in earlier studies (2) a starvation-like response could be observed in B. subtilis
401
upon sorbic acid stress.
402
With high concentrations of acetic acid, it is likely that fermentation routes towards acetate
403
are diverted to 2,3-butanediol. Synthesis of 2,3-butanediol depends on bdhA, encoding
404
acetoin reductase (31), a gene that has SigB-controlled expression. Acetic acid stress can
405
trigger stressosome activity and indeed results in up regulation of acetoin reductase
406
(Unpublished results by A. Ter Beek, (52)). Therefore, when acetate is present as a
407
preservative, putative organoleptic properties of 2,3-butanediol should be considered when B.
408
subtilis or related species may be present. Weak organic acid stress also causes an up
409
regulation of citric acid cycle enzymes (2, 52). This route may also consume added acetic
410
acid, thereby providing an extra carbon source and eliminating this weak organic acid. The
411
high oxygen consumption rate that we observed in the presence of 25 mM and 80 mM K-Ac,
412
may be explained by the utilization of acetate. While growth rates are reduced by 80% with
413
both 11 mM K-sorbate and 80 mM K-acetate, the glucose consumption rate with 80 mM K-
18
414
acetate is almost half of that with 11 mM K-sorbate. Future experiments should be conducted
415
in a turbidostat setup and include CO2 measurements to allow a direct measurement of the
416
carbon balance under these conditions. Therefore, knowledge of the pHi can be used to
417
predict glycolysis and this knowledge might be useful for modelling purposes in food
418
fermentation (47, 53).
419
In summary, oxidative phosphorylation is an important source of ATP for aerobic, non-
420
stressed B. subtilis cells. The O2 consumption rates follow a similar trend as qGlucose,
421
qAcetate and q2,3-Butanediol, apart from the 25 mM K-Ac stress. The level of WOA stress
422
may leave too little PMF to allow F1F0-ATPase to produce ATP. Sorbic acid has a similar
423
effect as acetic acid on pHi, but is more effective in depleting Δψ. This may partially explain
424
the observation that sorbic acid is a more potent preservative than acetic acid. In the quest for
425
preservatives active at near neutral pH, industry should best focus on ‘uncoupler-like’
426
molecules that act similar to or better than sorbic acid.
427
428
Acknowledgements
429
Jos Arents is thanked for his assistance with the membrane potential measurements. Gertien
430
Smits is acknowledged for critically reading initial versions of the manuscript.
431
432
433
434
435
436
437
438
19
439
References
440
441
1.
Eklund T. 1983. The antimicrobial effect of dissociated and undissociated sorbic acid
at different pH levels. J. Appl. Bacteriol. 54:383–9.
442
443
444
445
2.
Ter Beek A, Keijser BJF, Boorsma A, Zakrzewska A, Orij R, Smits GJ, Brul S.
2008. Transcriptome analysis of sorbic acid-stressed Bacillus subtilis reveals a nutrient
limitation response and indicates plasma membrane remodeling. J. Bacteriol.
190:1751–61.
446
447
3.
Stratford M, Anslow PA. 1998. Evidence that sorbic acid does not inhibit yeast as a
classic “weak acid preservative”. Lett. Appl. Microbiol. 27:203–6.
448
4.
International Programme on Chemical Safety (http://www.inchem.org/).
449
450
451
452
5.
Stratford M, Plumridge A, Nebe-von-Caron G, Archer DB. 2009. Inhibition of
spoilage mould conidia by acetic acid and sorbic acid involves different modes of
action, requiring modification of the classical weak-acid theory. Int. J. Food Microbiol.
136:37–43.
453
454
455
6.
Ullah A, Orij R, Brul S, Smits GJ. 2012. Quantitative analysis of the modes of
growth inhibition by weak organic acids in yeast. Appl. Environ. Microbiol. 78:8377–
87.
456
457
7.
Russell JB. 1992. Another explanation for the toxicity of fermentation acids at low pH:
anion accumulation versus uncoupling. J. Appl. Microbiol. 73:363–370.
458
459
460
8.
Carpenter CE, Broadbent JR. 2008. External concentration of organic acid anions
and pH: key independent variables for studying how organic acids inhibit growth of
bacteria in mildly acidic foods. J. Food Sci. 74:R12–5.
461
462
9.
Henderson L. 1908. Concerning the Relationship Between the Strength of Acids and
Their Capacity to Preserve Neutrality. Am. J. Physiol. 21:173–179.
463
464
10.
York GK, Vaughn RH. 1964. Mechanisms in the inhibition of microorganisms by
sorbic acid. J. Bacteriol. 88:411–7.
465
466
467
468
11.
Mollapour M, Fong D, Balakrishnan K, Harris N, Thompson S, Schüller C,
Kuchler K, Piper PW. 2004. Screening the yeast deletant mutant collection for
hypersensitivity and hyper-resistance to sorbate, a weak organic acid food preservative.
Yeast 21:927–46.
469
470
471
12.
Golden DA, Beuchat LR, Hitchcock HL. 1994. Changes in fatty acid composition of
various lipid components of Zygosaccharomyces rouxii as influenced by solutes,
potassium sorbate and incubation temperature. Int. J. Food Microbiol. 21:293–303.
472
473
474
13.
Casas E, de Ancos B, Valderrama MJ, Cano P, Peinado JM. 2004. Pentadiene
production from potassium sorbate by osmotolerant yeasts. Int. J. Food Microbiol.
94:93–6.
20
475
476
477
478
14.
Plumridge A, Stratford M, Lowe KC, Archer DB. 2008. The weak-acid
preservative sorbic acid is decarboxylated and detoxified by a phenylacrylic acid
decarboxylase, PadA1, in the spoilage mold Aspergillus niger. Appl. Environ.
Microbiol. 74:550–2.
479
480
15.
Mols M, Abee T. 2011. Bacillus cereus responses to acid stress. Environ. Microbiol.
13:2835–43.
481
482
16.
Brul S, Coote P. 1999. Preservative agents in foods. Mode of action and microbial
resistance mechanisms. Int. J. Food Microbiol. 50:1–17.
483
484
485
17.
Eklund T. 1985. The effect of sorbic acid and esters of p-hydroxybenzoic acid on the
protonmotive force in Escherichia coli membrane vesicles. J. Gen. Microbiol. 131:73–
6.
486
487
488
18.
Oomes SJCM, van Zuijlen ACM, Hehenkamp JO, Witsenboer H, van der Vossen
JMBM, Brul S. 2007. The characterisation of Bacillus spores occurring in the
manufacturing of (low acid) canned products. Int. J. Food Microbiol. 120:85–94.
489
490
491
19.
Keijser BJF, Ter Beek A, Rauwerda H, Schuren F, Montijn R, van der Spek H,
Brul S. 2007. Analysis of temporal gene expression during Bacillus subtilis spore
germination and outgrowth. J. Bacteriol. 189:3624–34.
492
493
494
20.
Van Beilen JWA, Brul S. 2013. Compartment-specific pH monitoring in Bacillus
subtilis using fluorescent sensor proteins; a tool to analyse the antibacterial effect of
weak organic acids. Front. Microb. Physiol. Metab.
495
496
497
21.
Lolkema J, Hellingwerf K, Konings W. 1982. The effect of “probe binding”on the
quantitative determination of the proton-motive force in bacteria. Biochim. Biophys.
Acta 150:1183–91.
498
499
500
501
22.
Lolkema JS, Abbing A, Hellingwerf KJ, Konings WN. 1983. The Transmembrane
Electrical Potential in Rhodopseudomonas sphaeroides Determined from the
Distribution of Tetraphenylphosphonium after Correction for Its Binding to Cell
Components. Eur. J. Biochem. 292:287–292.
502
503
504
23.
Zaritsky A, Kihara M, Macnab RM. 1981. Measurement of membrane potential in
Bacillus subtilis: A comparison of lipophilic cations, rubidium ion, and a cyanine dye
as probes. J. Membr. Biol. 63:215–231.
505
506
507
24.
De Vrij W, Driessen AJ, Hellingwerf KJ, Konings WN. 1986. Measurements of the
proton motive force generated by cytochrome c oxidase from Bacillus subtilis in
proteoliposomes and membrane vesicles. Eur. J. Biochem. 156:431–40.
508
509
25.
Salmond C V, Kroll RG, Booth IR. 1984. The effect of food preservatives on pH
homeostasis in Escherichia coli. J. Gen. Microbiol. 130:2845–50.
510
511
512
26.
Slonczewski JL, Fujisawa M, Dopson M, Krulwich TA. 2009. Cytoplasmic pH
measurement and homeostasis in bacteria and archaea. Adv. Microb. Physiol. 55:1–79,
317.
21
513
514
27.
Mitchell P. 1961. Coupling of phosphorylation to electron and hydrogen transfer by a
chemi-osmotic type of mechanism. Nature 191:144–8.
515
516
28.
Mitchell P. 1977. Vectorial chemiosmotic processes. Annu. Rev. Biochem. 46:996–
1005.
517
518
519
520
29.
Bulthuis BA, Koningstein GM, Stouthamer AH, van Verseveld HW. 1993. The
relation of proton motive force, adenylate energy charge and phosphorylation potential
to the specific growth rate and efficiency of energy transduction in Bacillus
licheniformis under aerobic growth conditions. Antonie Van Leeuwenhoek 63:1–16.
521
522
30.
Shioi JI, Matsuura S, Imae Y. 1980. Quantitative measurements of proton motive
force and motility in Bacillus subtilis. J. Bacteriol. 144:891–7.
523
524
525
31.
Cruz Ramos H, Hoffmann T, Marino M, Nedjari H, Presecan-Siedel E, Dreesen
O, Glaser P, Jahn D. 2000. Fermentative metabolism of Bacillus subtilis: physiology
and regulation of gene expression. J. Bacteriol. 182:3072–80.
526
527
528
32.
Frädrich C, March A, Fiege K, Hartmann A, Jahn D, Härtig E. 2012. The
transcription factor AlsR binds and regulates the promoter of the alsSD operon
responsible for acetoin formation in Bacillus subtilis. J. Bacteriol. 194:1100–12.
529
530
33.
Tännler S, Decasper S, Sauer U. 2008. Maintenance metabolism and carbon fluxes in
Bacillus species. Microb. Cell Fact. 7:19.
531
532
533
34.
Pickett MJ, Clifton CE. 1943. The effect of selective poisons on the utilization of
glucose and intermediate compounds by microorganisms. J. Cell. Comp. Physiol.
22:147–165.
534
535
536
35.
Stratford M, Anslow PA. 1996. Comparison of the inhibitory action on
Saccharomyces cerevisiae of weak-acid preservatives, uncouplers, and medium-chain
fatty acids. FEMS Microbiol. Lett. 142:53–8.
537
538
36.
Kinderlerer JL, Hatton P V. 1990. Fungal metabolites of sorbic acid. Food Addit.
Contam. 7:657–69.
539
540
37.
Russell JB, Diez-Gonzalez F. 1998. The effects of fermentation acids on bacterial
growth. Adv. Microb. Physiol. 39:205–34.
541
542
38.
Michels M, Bakker EP. 1981. The mechanism of uncoupling by picrate in
Escherichia coli K-12 membrane systems. Eur. J. Biochem. 116:513–9.
543
544
39.
Sikkema J, de Bont JA, Poolman B. 1995. Mechanisms of membrane toxicity of
hydrocarbons. Microbiol. Rev. 59:201–22.
545
546
40.
Leão C, Van Uden N. 1984. Effects of ethanol and other alkanols on passive proton
influx in the yeast Saccharomyces cerevisiae. Biochim. Biophys. Acta 774:43–8.
22
547
548
549
41.
Spycher S, Smejtek P, Netzeva TI, Escher BI. 2008. Toward a class-independent
quantitative structure--activity relationship model for uncouplers of oxidative
phosphorylation. Chem. Res. Toxicol. 21:911–27.
550
551
552
42.
Chu S, Hawes JW, Lorigan GA. 2009. Solid-state NMR spectroscopic studies on the
interaction of sorbic acid with phospholipid membranes at different pH levels. Magn.
Reson. Chem. 47:651–7.
553
554
555
556
557
43.
Holyoak CD, Stratford M, McMullin Z, Cole MB, Crimmins K, Brown AJ, Coote
PJ. 1996. Activity of the plasma membrane H(+)-ATPase and optimal glycolytic flux
are required for rapid adaptation and growth of Saccharomyces cerevisiae in the
presence of the weak-acid preservative sorbic acid. Appl. Environ. Microbiol.
62:3158–64.
558
559
44.
Pampulha ME, Loureiro-Dias MC. 2000. Energetics of the effect of acetic acid on
growth of Saccharomyces cerevisiae. FEMS Microbiol. Lett. 184:69–72.
560
561
562
45.
Ugurbil K, Rottenberg H, Glynn P, Shulman RG. 1978. 31P nuclear magnetic
resonance studies of bioenergetics and glycolysis in anaerobic Escherichia coli cells.
Proc. Natl. Acad. Sci. U. S. A. 75:2244–8.
563
564
46.
Krebs HA, Wiggins D, Stubbs M, Sols A, Bedoya F. 1983. Studies on the
mechanism of the antifungal action of benzoate. Biochem. J. 214:657–63.
565
566
567
47.
Vojinović V, von Stockar U. 2009. Influence of uncertainties in pH, pMg, activity
coefficients, metabolite concentrations, and other factors on the analysis of the
thermodynamic feasibility of metabolic pathways. Biotechnol. Bioeng. 103:780–95.
568
569
570
48.
Wu TF, Davis EJ. 1981. Regulation of glycolytic flux in an energetically controlled
cell-free system: the effects of adenine nucleotide ratios, inorganic phosphate, pH, and
citrate. Arch. Biochem. Biophys. 209:85–99.
571
572
573
49.
Kuhn N, Setlow B, Setlow P. 1995. Cooperative manganese (II) activation of 3phosphoglycerate mutase of Bacillus megaterium: a biological pH-sensing mechanism
in bacterial spore formation and germination. Arch. Biochem. Biophys. 320:35–42.
574
575
50.
Fujita Y. 2009. Carbon Catabolite Control of the Metabolic Network in Bacillus
subtilis. Biosci. Biotechnol. Biochem. 73:245–259.
576
577
51.
Sonenshein AL. 2007. Control of key metabolic intersections in Bacillus subtilis. Nat.
Rev. Microbiol. 5:917–27.
578
579
52.
Ter Beek AS. 2009. Weak organic acid stress in Bacillus subtilis. University of
Amsterdam.
580
581
582
53.
Van Heerden JH, Wortel MT, Bruggeman FJ, Heijnen JJ, Bollen YJM, Planqué
R, Hulshof J, O’Toole TG, Wahl SA, Teusink B. 2014. Lost in transition: start-up of
glycolysis yields subpopulations of nongrowing cells. Science 343:1245114.
23
583
584
585
54.
Culham DE, Romantsov T, Wood JM. 2008. Roles of K+, H+, H2O, and DeltaPsi in
solute transport mediated by major facilitator superfamily members ProP and LacY.
Biochemistry 47:8176–85.
586
587
55.
Strahl H, Hamoen LW. 2010. Membrane potential is important for bacterial cell
division. Proc. Natl. Acad. Sci. U. S. A. 107:12281–6.
588
24
589
Lengends to the figures:
590
591
Figure 1. Growth curve (A) and internal pH (B) of B. subtilis PB2 pDG-IpHluorin monitored
592
for 6 h under various stress conditions. Medium pH = 6.4. In M3G, the growth rate is highest,
593
but a continuous decrease in pHi is seen. Sorbic acid stress causes the growth rate to be
594
reduced as well as a continuous decrease in internal pH. Acetic acid stress has a similar effect
595
on growth rate, but pHi remains constant.
596
597
Figure 2. Acidification and recovery upon weak organic acid addition to B. subtilis PB2
598
PptsG-IpHluorin. Sorbic acid (A and B) or acetic acid (C and D) is injected at t = 0 s. Medium
599
pH = 6.4. The internal pH drops to its lowest point within seconds (A and C) and recovers
600
fast when glucose is available (B and D). The pH recovers to higher values with acetic acid
601
stress. Data from typical examples is shown.
602
603
Figure 3. Acidification and recovery upon weak organic acid addition to B. subtilis PB2
604
PptsG-IpHluorin. Acetic or sorbic acid is injected at t = 10s. Medium pH = 6.4. The internal
605
pH drops to its lowest point within seconds and recovers fast when glucose is available. With
606
glucose available, the pHi at t = 0 is also higher. The pH recovers to higher values with acetic
607
acid stress. Data from a typical example is shown
608
609
Figure 4. Amplitude of acidification by addition of sorbic or acetic acid. A linear trend can
610
be observed between ΔpH and –log[HA] ([HA] in M). Data from experiments at an
611
extracellular pH=5.5 and 6.4 are combined. Error bars indicate standard deviation.
612
613
25
614
Figure 5. Differential effects of sorbic and acetic acid on the proton motive force. Sorbic acid
615
stress affects both ΔpH as well as Δψ. Acetic acid affects ΔpH, but has little effect on Δψ.
616
Data are from cultures exposed to weak acids for approximately 5 min. Results are based on
617
3 biological replicates. Error bars indicate standard deviation.
618
619
Figure 6. Weak organic acid stress reduces respiration. Respiration was monitored for 1 min
620
or the time it took to reduce O2 levels to 0 µmol. Results are based on 3 biological replicates.
621
Error bars indicate standard deviation.
622
623
Figure 7. Increased weak organic acid stress lowers glucose consumption rate. The rate of
624
glucose consumption compared to the rate of acetate and 2,3-butanediol production. Results
625
are based on two biological replicates. Error bars indicate standard deviation.
626
627
Figure 8. Glucose flux versus internal pH due to weak organic acid stress (no stress, 3 and 11
628
mM K-S, 25 and 80 mM K-Ac). A decrease in pHi due to WOA stress is accompanied by a
629
decrease in glucose flux.
630
26