Download Application of the new manP counter-selection system for B. subtilis

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Neuronal ceroid lipofuscinosis wikipedia , lookup

Genetically modified crops wikipedia , lookup

Essential gene wikipedia , lookup

Extrachromosomal DNA wikipedia , lookup

Copy-number variation wikipedia , lookup

Polycomb Group Proteins and Cancer wikipedia , lookup

Transposable element wikipedia , lookup

Saethre–Chotzen syndrome wikipedia , lookup

Human genome wikipedia , lookup

Non-coding DNA wikipedia , lookup

NEDD9 wikipedia , lookup

Oncogenomics wikipedia , lookup

Molecular cloning wikipedia , lookup

Genomic imprinting wikipedia , lookup

X-inactivation wikipedia , lookup

Gene therapy of the human retina wikipedia , lookup

Epigenetics of diabetes Type 2 wikipedia , lookup

RNA-Seq wikipedia , lookup

Genomics wikipedia , lookup

Epigenetics of human development wikipedia , lookup

Point mutation wikipedia , lookup

Minimal genome wikipedia , lookup

Cre-Lox recombination wikipedia , lookup

Gene nomenclature wikipedia , lookup

Public health genomics wikipedia , lookup

Plasmid wikipedia , lookup

Gene desert wikipedia , lookup

Gene therapy wikipedia , lookup

Nutriepigenomics wikipedia , lookup

Gene expression programming wikipedia , lookup

Pathogenomics wikipedia , lookup

Gene wikipedia , lookup

Gene expression profiling wikipedia , lookup

Therapeutic gene modulation wikipedia , lookup

Vectors in gene therapy wikipedia , lookup

Genetic engineering wikipedia , lookup

Genome editing wikipedia , lookup

Helitron (biology) wikipedia , lookup

Genome evolution wikipedia , lookup

Genome (book) wikipedia , lookup

Genomic library wikipedia , lookup

Microevolution wikipedia , lookup

No-SCAR (Scarless Cas9 Assisted Recombineering) Genome Editing wikipedia , lookup

History of genetic engineering wikipedia , lookup

Designer baby wikipedia , lookup

Artificial gene synthesis wikipedia , lookup

Site-specific recombinase technology wikipedia , lookup

Transcript
1
Development of a markerless gene deletion system for Bacillus subtilis based
2
on the mannose phosphoenolpyruvate-dependent phosphotransferase system
3
4
Marian Wenzel and Josef Altenbuchner*
5
6
Institut für Industrielle Genetik, Universität Stuttgart, Allmandring 31, 70569 Stuttgart,
7
Germany
8
9

Corresponding author: Josef Altenbuchner
10
E-mail: [email protected]
11
Tel. ++49 711 685 67591
12
Fax: ++49 711 685 66973
13
14
Running title:
15
Markerless gene deletion system for Bacillus subtilis
16
17
Keywords:
18
PTS, genome engineering, counter-selection marker, mannose degradation
19
20
Abstract
21
To optimize Bacillus subtilis as a production strain for proteins and low molecular
22
substances by genome engineering, we developed a markerless gene deletion
23
system. We took advantage of a general property of the phosphoenolpyruvate-
24
dependent phosphotransferase system (PTS), in particular the mannose PTS.
25
Mannose is phosphorylated during uptake by its specific transporter (ManP) to
26
mannose 6-phosphate which is further converted to fructose 6-phosphate by the
27
mannose 6-phosphate isomerase (ManA). When ManA is missing, accumulation of
28
the phosphorylated mannose inhibits cell growth. This system was realized by
29
deletion of manP and manA in B. subtilis 6, a 168 derivative strain with 6 large
30
deletions of prophages and antibiotic biosynthesis genes. The manP gene was
31
inserted into an E. coli plasmid together with a spectinomycin resistance gene for
32
selection in B. subtilis. To delete a specific region, its up- and downstream flanking
33
sites (each approx. 700 bps) were inserted into the vector. After transformation,
34
integration of the plasmid into the chromosome of B. subtilis by single cross-over was
35
selected by spectinomycin. In the second step, excision of the plasmid was selected
36
by growth on mannose. Finally, excision and concomitant deletion of the target
37
region was verified by Colony PCR. In this way, all 9 prophages, 7 antibiotic
38
biosynthesis gene clusters, 2 sigma factors for sporulation were deleted and the
39
B. subtilis genome was reduced from 4,215 to 3,640 kb. Despite these extensive
40
deletions, growth rate and cell morphology remained similar to the B. subtilis 168
41
parental strain.
42
43
Introduction
44
Bacillus subtilis is an important producer of industrial enzymes, such as amylases
45
and proteases, with favorable features like high protein secretion capability, GRAS
46
(generally recognized as safe) status, natural competence, a long tradition in
47
fermentation and a deep knowledge of its genetics, physiology and biochemistry. Up-
48
to-date information is available on SubtiWiki (Michna et al., 2014). Strain optimization
49
to remove adverse properties like sporulation, lysogenic prophages, production of
50
antibiotics and extracellular proteases (Krishnappa et al., 2013; Stein, 2005; Westers
51
et al., 2003; Westers et al., 2004) or to introduce new metabolic routes, requires
52
powerful tools for genetic engineering.
53
An easy way for chromosomal deletions is to construct deletion cassettes comprising
54
an antibiotic resistance marker flanked by sequences that normally flank the region to
55
be deleted. The successful integration of the cassette via homologous recombination
56
is selected by the antibiotic resistance. For multiple modifications of the
57
chromosome, recognition sites for site-specific recombinases, such as Xer / dif, Flp /
58
FRT or Cre / loxP (Bloor & Cranenburgh, 2006; Hoang et al., 1998; Marx & Lidstrom,
59
2002), allow the precise excision of the marker gene, and the cassette can be used
60
again. The risk of unintended large chromosomal deletions or inversions can be
61
avoided by using chimeric recognition sites like lox71/lox66 or mroxP (Warth &
62
Altenbuchner, 2013a; Warth & Altenbuchner, 2013b). However, a second
63
transformation and curing of the plasmid providing the recombinase is needed. Last
64
but not least, all such methods leave ‘scars’ in the chromosome.
65
The development of markerless systems deals with this problem. Here, the antibiotic
66
resistance gene is located outside the cloned flanking sequences on a suicide vector.
67
The vector is integrated via single cross-over recombination. Spontaneous excision
68
of the vector either leads to removal of just the vector or to removal of the vector and
69
the target region. However, this is a rare event. By integration of a -galactosidase
70
gene into the vector it is possible to detect such rare events by blue-white screening
71
but this needs strains lacking -galactosidase activity and the screening of thousands
72
of colonies (Leenhouts et al., 1996). Therefore, counter-selection marker systems
73
have been developed, like the popular fusaric acid (tetAR), streptomycin (rpsL), and
74
sucrose (sacB) sensitivity systems (reviewed by Reyrat et al., 1998). Another one
75
takes advantage of the endonuclease I-SceI of Saccharomyces cerevisiae (Martínez-
76
Garcia & de Lorenzo, 2011; Monteilhet et al., 1990). However, the application of
77
these systems has not been proven effective or even applicable in B. subtilis due to
78
natural resistances, lack of suitable vectors etc.
79
A rather powerful system based on the sensitivity against 5-fluorouracil (5-FU; upp)
80
has been developed for B. subtilis and successfully adapted for several other
81
microorganisms (Fabret et al., 2002; Goh et al., 2009; Graf & Altenbuchner, 2011;
82
Keller et al., 2009; Kristich et al., 2005; Tanaka et al., 2013). Besides the toxicity of 5-
83
FU, a major disadvantage is the occurrence of diverse suppressor mutations as
84
already stated by Fabret et al. (2002). Other systems based on the action of toxins
85
and antitoxins such as CcdB and MazF from E. coli (Bernard et al., 1994; Zhang et
86
al., 2006) require a perfect and sensitive toxin/antitoxin equilibrium. As these systems
87
deal with toxic molecules, the occurrence of suppressor mutations is also very likely.
88
Alternatively, temperature-sensitive plasmids allow selection of their integration at
89
their non-permissive temperature. Returning to its permissive temperature, the
90
plasmid starts replicating again, and excision is selected (Arnaud et al., 2004;
91
Zakataeva et al., 2010). However, the cells might also survive by slow growth and
92
one has to make sure that the plasmid was indeed integrated.
93
Thus, the need for a less toxic, but powerful and reliable counter-selection system for
94
B. subtilis in order to deal with the upcoming questions from large scale omics-data
95
and analyses is obvious and necessary (Buescher et al., 2012; Goelzer et al., 2008;
96
Kobayashi et al., 2001; Petersohn et al., 2001; Price et al., 2001; Yoshida et al.,
97
2001). In this study, we describe the development and application of a novel,
98
markerless method for multiple subsequent gene deletions in B. subtilis. We took
99
advantage of a general property of the phosphoenolpyruvate-dependent
100
phosphotransferase system (PTS), and particularly the mannose PTS. With this
101
method we were able to construct a strain without several adverse properties and a
102
reduced genome. The resulting strain was examined concerning its growth rate.
103
104
Materials and Methods
105
Plasmids, bacterial strains and growth conditions
106
Relevant bacterial strains and plasmids used in this study are summarized in Table 1
107
and Table S1. Standard recombinant DNA techniques were used (Sambrook et al.,
108
1989). Cloning steps were performed with E. coli JM109 (Yanish-Perron et al., 1985).
109
E. coli JM109 was transformed with plasmid DNA using the TSS heat shock method
110
as described before (Chung et al., 1989). B. subtilis was transformed according to
111
the modified “Paris method” (Harwood & Cutting, 1990). All strains were grown at
112
37°C, using LB medium (Bertani, 1951) or MG1. The MG1 minimal medium used for
113
the first of the two-step growth phase for competence development and for
114
comparing the growth rates of B. subtilis contained per 1 liter: 2 g (NH4)2SO4, 6 g
115
KH2PO4, 14 g K2HPO4, 1 g Na-citrate, 0.2 g MgSO4, 5 g glucose, 200 mg casamino
116
acids, 50 mg tryptophan (if necessary) and 1 ml of a 1000-fold trace element solution
117
(TES; Wenzel et al., 2011). Antibiotics were used in the following concentrations:
118
ampicillin (amp), 100 µg ml-1; spectinomycin (spc), 100 µg ml-1; erythromycin (erm) 5
119
µg ml-1.
120
Materials
121
Mannose was purchased from AMRESCO LLC (Solon, Ohio, USA). DNA modifying
122
enzymes including restriction endonucleases were purchased from New England
123
Biolabs GmbH (Frankfurt am Main, Germany). PCRs were run with High Fidelity PCR
124
Enzyme Mix and DreamTaq DNA polymerase (Thermo Fisher Scientific GmbH,
125
Schwerte, Germany) on a TPersonal Thermocycler (Biometra GmbH, Göttingen,
126
Gemany). Synthetic DNA oligonucleotides (Table S2) were purchased from Eurofins
127
MWG Operon GmbH (Ebersberg, Germany). DNA sequencing was performed by
128
GATC Biotech AG (Constance, Germany).
129
Construction of plasmids pJOE6743.1 and pJOE6577.1
130
The plasmid pJOE6743.1 is derived from pJOE4786.1 (Jeske & Altenbuchner, 2010).
131
The spectinomycin resistance gene was isolated from pMW168.1 (Wenzel et al.,
132
2011) by endoR AgeI and EcoRI, the protruding ends of the fragment filled in by
133
Klenow polymerase treatment and inserted into the SfoI site of pJOE4786.1
134
(pJOE6562.4). The manP gene together with its promoter was amplified from B.
135
subtilis 168 by PCR using primers s6477 and s6041, cleaved with NdeI and inserted
136
into the NdeI site of pJOE6562.4 to give pJOE6743.1. The plasmid pJOE6577.1 is
137
derived from pSUN346.1 (Sun & Altenbuchner, 2010). The plasmid pSUN346.1
138
again is derived from pJOE4786.1 and contains the C-terminal end of manR, the
139
manA gene, followed by an erythromycin resistance gene (erm), ydjF and a
140
spectinomycin resistance gene (aad9). The manA gene was removed by cutting
141
pSUN346.1 with BamHI and PmeI and after Klenow treatment ligated to plasmid
142
pJOE6577.1.
143
Protocol for the chromosomal deletion formation
144
Derivatives of pJOE6743.1 with deletion cassettes were used to transform a manPA
145
negative B. subtilis strain. Colonies with the integrated plasmid (via single cross-over)
146
were selected on LB agar plates containing spectinomycin. Five of the transformants
147
were picked and streaked on a new LB agar plate containing spectinomycin. One
148
single colony obtained from each of the five transformants was used to inoculate LB
149
liquid medium. After incubation at 37°C for about 8 h the cultures were diluted 10-4 in
150
LB liquid medium containing 0.5% (w/v) mannose. After incubation at 37°C overnight,
151
10-6 dilutions were plated on LB agar plates containing 0.5% (w/v) mannose. The
152
mannose in the last two steps helps to select the cells which have lost the integrated
153
vector (counter-selection). Four colonies from each plate (20 clones at total) were
154
picked and streaked on LB agar plates without and with spectinomycin.
155
Spectinomycin sensitive colonies were finally checked by Colony PCR.
156
Colony PCR
157
After selection for chromosomal plasmid excision, colonies were streaked out on LB
158
agar plates. A small amount of cells was taken from these plates and resuspended in
159
100 µl H2O (bidest.). After heating the suspension for 10 min at 99°C, a cold shock
160
for 20 min at -70°C followed by a further heating step (10 min at 99°C) was
161
performed. After centrifugation of the cell suspension, 2 µl of the supernatant
162
containing chromosomal DNA was used in a 30 µl PCR preparation using the
163
DreamTaq DNA polymerase. As a control, the same procedure was performed with
164
two colonies still containing the integrated plasmid.
165
166
Results
167
Plasmid and strain construction for manP-based counter-selection
168
In order to develop a new efficient counter-selection system for B. subtilis, we took
169
advantage of the fact that mannose 6-phosphate dehydrogenase mutants (manA)
170
were sensitive to the presence of mannose (Sun & Altenbuchner, 2010; Wenzel et
171
al., 2011). Cells developed a bubble like morphology and stopped growing about 1 h
172
after addition of mannose to the culture medium (Wenzel et al., 2011). The sensitivity
173
against mannose is dependent on the presence of manP and absence of manA. A
174
strain without both genes is resistant to mannose. The introduction of manP to such a
175
strain would reestablish the sensitivity against mannose. Taking advantage of these
176
observations, we developed the idea of using manP as a counter-selectable marker.
177
To get this to work one has to delete the manPA genes in the B. subtilis chromosome
178
and to develop a vector which i) carries the manP gene, ii) does not replicate in B.
179
subtilis, iii) encodes an antibiotic resistance gene for selection in B. subtilis and iiii)
180
has restriction sites for cloning of the deletion cassettes. The deletion cassettes, two
181
sequences of about 700 bp flanking the target region, enable the integration of the
182
vector into the B. subtilis chromosome via single cross-over. This is selected by the
183
antibiotic resistance. After the integration, cells are cultured in the presence of
184
mannose. Only cells which spontaneously have lost the vector, alone or together with
185
the target region, form colonies on mannose-containing agar plates. The colonies are
186
screened for loss of the antibiotic resistance and the antibiotic sensitive ones finally
187
tested by Colony PCR for loss of the target region.
188
To use manP for multiple deletions in the chromosome of B. subtilis the vector
189
pJOE6743.1 was constructed (Fig. 1). This is a cloning vector derived from
190
pJOE4786.1 (Jeske & Altenbuchner, 2010). Two long inverted repeats separated by
191
a lacPOZ fragment allow a blue/white screening in strains like E. coli JM109. When
192
the two repeats come together, the plasmid cannot be replicated (Altenbuchner et al.,
193
1992). Therefore, a positive selection occurs on the replacement of the lacPOZ
194
fragment. The vector is a pUC derivative and carries the ampicillin resistance gene
195
for selection in E. coli, a spectinomycin resistance gene for selection in both B.
196
subtilis and E. coli as well as the B. subtilis manP gene encoding the mannose PTS
197
transporter under control of its own promoter regulated by ManR (Wenzel &
198
Altenbuchner, 2013).
199
Before this vector can be used in B. subtilis, the manPA genes had to be removed. B.
200
subtilis 6, a derivative of B. subtilis 168 with a series of already deleted prophages
201
and antibiotic biosynthesis genes was chosen as the starting strain (Westers et al.,
202
2003). For an easier handling, the tryptophan auxotrophic strain was transformed
203
with chromosomal DNA from B. subtilis 3NA (Michel & Millet, 1970) to replace the
204
trpC2 allele present in the B. subtilis 168 derivatives (Table 1; for more details see
205
Table S1). To make sure that the new strain designated IIG-Bs1 still contained the
206
deletions conducted by Westers et al. (2003) PCRs were performed with primers
207
shown in Table S2.
208
Next, the manPA genes had to be deleted in the IIG-Bs1 chromosome. This was
209
done by cloning the upstream mannose regulatory gene and the downstream yjdF
210
gene to both sides of an erythromycin resistance gene in an E. coli pJOE4786.1
211
derivative. This plasmid had in addition a spectinomycin resistance gene inserted
212
downstream of yjdF (plasmid pJOE6577.1). This plasmid-DNA was used to transform
213
B. subtilis IIG-Bs1. The resulting strain designated IIG-Bs2 had the manPA genes
214
replaced by an erythromycin resistance gene and served as the initial strain for
215
further deletions using the new counter-selection system. Later, the erythromycin
216
resistance gene was deleted together with yjdF and several further genes
217
downstream of yjdF (IIG-Bs20-1, Table 1 and Table S1).
218
Development and application of the new counter-selection method
219
First, the flanking regions of the target DNA to be deleted were amplified by PCR.
220
The fragments were combined and cloned into the vector pJOE6743.1 in three
221
different ways: i) restriction sites were added by the primers to both sides of the
222
amplified fragments and the fragments cloned into the vector by a three-fragment
223
ligation reaction; ii) amplified fragments were fused by overlap extension PCR using
224
the outer primers (Ho et al., 1989) and inserted blunt ended into the SmaI or EcoRV
225
cut vector; iii) restriction sites were added to the outer ends of the amplified
226
fragments, the fragments fused by overlap PCR, cut at the corresponding restriction
227
sites and inserted into the vector cleaved with the same restriction enzyme. Tables
228
S1 and S2 summarize the constructed plasmids and oligonucleotides used for PCR.
229
Using a pJOE6743.1 derivative with a deletion cassette, a manPA negative strain
230
was transformed. In principle, the integration of the plasmid in the chromosome was
231
selected by spectinomycin and in a second step, the excision was selected, first by
232
growing the cells in LB liquid, in LB liquid containing mannose and on mannose
233
containing LB plates. The excision and deletion formation was checked by Colony
234
PCR. In theory, half of the clones should have lost the region of interest, while the
235
other half should have restored the original state. In fact, we observed that on
236
average, 30 - 70 % of the clones had lost the region of interest.
237
Application of the new manP counter-selection system for B. subtilis genome
238
engineering
239
Ten large AT rich regions were identified in the B. subtilis 168 chromosome, that
240
represent the temperate phage SP (Zahler et al., 1977), the defective prophage
241
PBSX (Wood et al., 1990), the integrative conjugative element ICEBs1 (Auchtung et
242
al., 2005), skin (sigma K intervening) integrated in sigK and 6 prophage like elements
243
(Westers et al., 2003). Lysogenic phages and cryptic prophages might cause severe
244
problems by spontaneous or induced phage activation during fermentation.
245
Therefore, it is advantageous to remove all phage elements. In B. subtilis 6 the
246
phages SB and PBSX as well as the cryptic prophages pro1, pro3 and skin have
247
already been deleted. With the new counter-selection method described in the
248
section above we were able to successively delete ICEBs1 together with a large
249
flanking region (IIG-Bs3), pro5 (IIG-Bs4), pro6 (IIG-Bs5), pro4 (IIG-Bs20-2) and
250
pro7 (IIG-Bs12) (see Table 1 and Table S1).
251
About 4-5% of the B. subtilis genome is devoted to antibiotic production (Stein,
252
2005). There are two lantibiotic biosynthesis gene clusters producing sublancin 168
253
(Paik et al., 1998) and subtilosin A (Zheng et al., 2000). Two further gene clusters are
254
responsible for the production of the non-ribosomal lipopeptides surfactin and
255
plipastatin (synonymous to fengycin). The pks operon encodes the synthesis of an
256
unknown polyketide (Kunst et al., 1997). Due to a mutation in sfp, the
257
phosphopantetheine transferase, the non-ribosomal peptides and the polyketide are
258
not produced in B. subtilis 168 (Mootz et al., 2001). Nevertheless, the pathways can
259
be reactivated by simple transformation with DNA of B. subtilis wild type strains
260
(Nakano et al., 1992; Tsuge et al., 1999). There are three other antibiotics produced
261
by B. subtilis 168, the dipeptide bacilysin (Inaoka et al., 2003), the phospholipid
262
bacilysocin (Tamehiro et al., 2002) and the amino sugar 3,3´-neotrehalosadiamine
263
(Inaoka et al., 2004). The sublancin 168 gene cluster is located on phage SP and
264
was already deleted in 6. The pks operon was deleted as well in 6 and replaced by
265
a chloramphenicol resistance gene. To get rid of this antibiotic resistance gene, the
266
original deletion was slightly extended leading to IIG-Bs8. The subtilosin A production
267
genes were removed by deletion of genes sboA, sboX, albA-albG (IIG-Bs9) with the
268
new method.
269
In case of plipastatin the genes ppsA-ppsE (Steller et al., 1999) were deleted (IIG-
270
Bs10). In case of surfactin the situation was more complicated (Marahiel et al., 1993).
271
Within srfA, encoding the nonribosomal peptide synthase, there is a second small
272
gene comS needed for competence. (Turgay et al., 1998). Hence, srfAA, srfAB
273
(except comS), srfAC, srfAD, and additional genes up to sfp were deleted. The srf-
274
promoter and comS, however, were retained (IIG-Bs20-4).
275
The bacilysin production was abolished by deletion of the genes bacABCDEF (IIG-
276
Bs11), the bacilysocin synthesis by deletion of ytpA (IIG-Bs13) and finally the 3,3´-
277
neotrehalosadiamine biosynthesis by deletion of ntdR (yhjM), ntdA, ntdB, ntdC and
278
glcP (IIG-Bs20-3). In the end, all known antibiotic biosynthesis clusters on the B.
279
subtilis chromosome are deleted in strain IIG-Bs20-4.
280
Endospore formation in B. subtilis is triggered by nutrient limitations. The master
281
regulator for differentiation into spores is Spo0A which becomes activated by
282
phosphorylation (Fujita & Losick, 2005). Activated Spo0A triggers two parallel and
283
interconnected programs of gene expression, one in the forespore by activating
284
sigma factor F and one in the mother cell by activating sigma factor E
285
(Eichenberger et al., 2003; Molle et al., 2003; Wang et al., 2006). Deletion of spo0A
286
would be the simplest way to inhibit sporulation. On the other hand, this would
287
inactivate the natural transformation capability. Therefore, we deleted in one step
288
sigma factor sigE and sigG together with spoGA responsible for pro-E activation and
289
a gene for the extracellular protease Bpr (IIG-Bs20).
290
During transition from exponential growth to sporulation some differentiating cells
291
show a fratricidal behaviour (González-Pastor, 2011). They produce toxins which kill
292
sibling cells. There are two independent gene clusters for toxin production: skf for
293
sporulation killing factor and sdp for sporulation delaying protein (Allenby et al.,
294
2006). The skf gene cluster was already removed by deleting prophage 1. Similarly
295
to the skf operon, the sdp gene cluster produces a toxic peptide which is modified
296
and exported by enzymes encoded by genes in two convergent operons, sdpABC
297
and sdpRI (Ellermeier et al., 2006; Engelberg-Kulka et al., 2006). The two sdp
298
operons were deleted together in strain IIG-Bs14 in order to get rid of upcoming killer
299
cells.
300
Comparison of the growth characteristics of IIG-Bs20-4 with B. subtilis 168
301
The chromosomal deletions in IIG-Bs20-4 might affect the growth rate of this strain.
302
Therefore we compared the growth curves of IIG-Bs20-4 with the ones from B.
303
subtilis 168 in LB and minimal media. For LB, overnight cultures in LB at 37°C were
304
used to inoculate 10 ml fresh LB medium in shake flasks to an optical density at 600
305
nm (OD600) of 0.05. The cultures were incubated at 37°C on a rotary shaker at 150
306
rpm. The OD600 was determined hourly. For minimal media, MG1 medium
307
supplemented with trace elements was used. Fresh MG1 medium was inoculated by
308
an overnight culture in MG1 to an OD600 of 0.1 and the growth determined
309
spectroscopically like in LB.
310
Independent of the used medium, both strains behaved similarly (Fig. 2). Using LB or
311
MG1 the two strains showed nearly identical growth rates (Table 2). Thus, even a
312
deletion of as much as 13.6 % of the B. subtilis genome did not compromise the
313
viability of the bacteria suggesting that the constructed strain IIG-Bs20-4 is a suitable
314
chassis for industrial applications.
315
Discussion
316
A counter-selection system is essential for markerless genome engineering. The
317
method described in this study allows to introduce point mutations, deletions of
318
various length as well as gene insertions into the B. subtilis chromosome. There is a
319
shortage of reliable counter-selection systems, especially of ones which do not use
320
toxic compounds. The use of the manP PTS transporter in a manPA mutant turned
321
out to be a highly reproducible and efficient way to produce a series of deletions in
322
the B. subtilis genome. No spontaneous resistance to mannose, for example by
323
mutation of the mannose activator gene or by induction of a phosphosugar
324
phosphatase was observed. The deletions from B. subtilis 6 to IIG-Bs20-4 were just
325
checked by Colony PCR. Another strain derived from IIG-Bs20 with many more
326
deletions made for a different purpose with this new counter-selection method clearly
327
showed after sequencing that all deletions occurred exactly as planed and there were
328
no DNA rearrangements or single base mutations found (D. Reuss, pers.
329
communication). Since many bacteria use PTS systems for sugar uptake, it should
330
be easy to adapt such a system for genome engineering in other bacterial species.
331
In a series of deletions we were able to considerably reduce the B. subtilis genome
332
from 4,215 to 3,640 kb. Another B. subtilis 168 strain (MGB874) with reduced
333
genome was constructed before (Morimoto et al., 2008). In MGB874, depleted of 874
334
kb (20%) of its genome sequence, the purpose was to delete a large number of
335
dispensable genes without altering the growth rate and other strain-specific
336
properties like sporulation or antibiotic production. Since many of the dispensable
337
genes are located in regions identified as prophages or genomic islands, the strains
338
MGB874 and IIG-Bs20-4 have many deleted genes in common. Nevertheless, IIG-
339
Bs20-4 has the advantage that is does not produce spores, antibiotics or
340
bacteriocins, properties which are desired for a production strain for enzymes or low
341
molecular substances.
342
The new strain provides a valuable basis for further optimization concerning protein
343
production and secretion, high cell-density fermentation or metabolic engineering.
344
345
Acknowledgements
346
We thank Annette Schneck and Sina Blessing for their assistance in creating the
347
deletion strains and Kambiz H. Morabbi for helpful discussions. The project was
348
partly financed by Lonza AG and thus we would like to thank especially Christoph
349
Kiziak for support and helpful discussions. We are grateful to Jan Maarten van Dijl for
350
providing the strain B. subtilis 6.
351
352
References
353
354
355
356
Allenby, N. E., Watts, C. A., Homuth, G., Prágai, Z., Wipat, A., Ward, A. C. &
Harwood, C. R. (2006). Phosphate starvation induces the sporulation
killing factor of Bacillus subtilis. J Bacteriol 188, 5299-5303.
357
358
Altenbuchner, J., Viell, P. & Pelletier, I. (1992). Positive selection vectors based
on palindromic DNA sequences. Methods Enzymol 216, 457-466.
359
360
361
Arnaud, M., Chastanet, A. & Débarbouillé, M. (2004). New vector for efficient
allelic replacement in naturally nontransformable, low-GC-content, grampositive bacteria. Appl Environ Microbiol 70, 6887-6891.
362
363
364
365
Auchtung, J. M., Lee, C. A., Monson, R. E., Lehman, A. P. & Grossman, A. D.
(2005). Regulation of a Bacillus subtilis mobile genetic element by
intercellular signaling and the global DNA damage response. Proc Natl
Acad Sci U S A 102, 12554-12559.
366
367
Bernard, P., Gabant, P., Bahassi, E. M. & Couturier, M. (1994). Positive-selection
vectors using the F plasmid ccdB killer gene. Gene 148, 71-74.
368
369
Bertani, G. (1951). Studies on lysogenesis. I. The mode of phage liberation by
lysogenic Escherichia coli. J Bacteriol 62, 293-300.
370
371
372
Bloor, A. E. & Cranenburgh, R. M. (2006). An efficient method of selectable
marker gene excision by Xer recombination for gene replacement in
bacterial chromosomes. Appl Environ Microbiol 72, 2520-2525.
373
374
375
376
Buescher, J. M., Liebermeister, W., Jules, M., Uhr, M., Muntel, J., Botella, E.,
Hessling, B., Kleijn, R. J., Le Chat, L. & other authors (2012). Global
network reorganization during dynamic adaptations of Bacillus subtilis
metabolism. Science 335, 1099-1103.
377
378
379
Chung, C. T., Niemela, S. L. & Miller, R. H. (1989). One-step preparation of
competent Escherichia coli: transformation and storage of bacterial cells
in the same solution. Proc Natl Acad Sci U S A 86, 2172-2175.
380
381
382
383
Eichenberger, P., Jensen, S. T., Conlon, E. M., van Ooij, C., Silvaggi, J.,
Gonzalez-Pastor, J. E., Fujita, M., Ben Yehuda, S., Stragier, P. & other
authors (2003). The sigmaE regulon and the identification of additional
sporulation genes in Bacillus subtilis. J Mol Biol 327, 945-972.
384
385
386
Ellermeier, C. D., Hobbs, E. C., Gonzalez-Pastor, J. E. & Losick, R. (2006). A
three-protein signaling pathway governing immunity to a bacterial
cannibalism toxin. Cell 124, 549-559.
387
388
389
Engelberg-Kulka, H., Amitai, S., Kolodkin-Gal, I. & Hazan, R. (2006). Bacterial
programmed cell death and multicellular behavior in bacteria. PLoS
Genet 2, e135.
390
391
Fabret, C., Ehrlich, S. D. & Noirot, P. (2002). A new mutation delivery system for
genome-scale approaches in Bacillus subtilis. Mol Microbiol 46, 25-36.
392
393
394
Fujita, M. & Losick, R. (2005). Evidence that entry into sporulation in Bacillus
subtilis is governed by a gradual increase in the level and activity of the
master regulator Spo0A. Genes Dev 19, 2236-2244.
395
396
397
398
Goelzer, A., Bekkal, B. F., Martin-Verstraete, I., Noirot, P., Bessières, P.,
Aymerich, S. & Fromion, V. (2008). Reconstruction and analysis of the
genetic and metabolic regulatory networks of the central metabolism of
Bacillus subtilis. BMC Syst Biol 2, 20.
399
400
401
402
403
Goh, Y. J., Azcárate-Peril, M. A., O'Flaherty, S., Durmaz, E., Valence, F., Jardin,
J., Lortal, S. & Klaenhammer, T. R. (2009). Development and application
of a upp-based counterselective gene replacement system for the study
of the S-layer protein SlpX of Lactobacillus acidophilus NCFM. Appl
Environ Microbiol 75, 3093-3105.
404
405
González-Pastor, J. E. (2011). Cannibalism: a social behavior in sporulating
Bacillus subtilis. FEMS Microbiol Rev 35, 415-424.
406
407
408
Graf, N. & Altenbuchner, J. (2011). Development of a method for markerless
gene deletion in Pseudomonas putida. Appl Environ Microbiol 77, 55495552.
409
410
Harwood, C. R. & Cutting, S. M. (1990). Molecular Biological Methods for Bacillus.
Chichester, England: John Wiley & Sons Ltd.
411
412
413
Ho, S. N., Hunt, H. D., Horton, R. M., Pullen, J. K. & Pease, L. R. (1989). Sitedirected mutagenesis by overlap extension using the polymerase chain
reaction. Gene 77, 51-59.
414
415
416
417
418
Hoang, T. T., Karkhoff-Schweizer, R. R., Kutchma, A. J. & Schweizer, H. P.
(1998). A broad-host-range Flp-FRT recombination system for sitespecific excision of chromosomally-located DNA sequences: application
for isolation of unmarked Pseudomonas aeruginosa mutants. Gene 212,
77-86.
419
420
421
422
Inaoka, T., Takahashi, K., Ohnishi-Kameyama, M., Yoshida, M. & Ochi, K. (2003).
Guanine nucleotides guanosine 5'-diphosphate 3'-diphosphate and GTP
co-operatively regulate the production of an antibiotic bacilysin in
Bacillus subtilis. J Biol Chem 278, 2169-2176.
423
424
425
426
Inaoka, T., Takahashi, K., Yada, H., Yoshida, M. & Ochi, K. (2004). RNA
polymerase mutation activates the production of a dormant antibiotic
3,3'-neotrehalosadiamine via an autoinduction mechanism in Bacillus
subtilis. J Biol Chem 279, 3885-3892.
427
428
429
Jeske, M. & Altenbuchner, J. (2010). The Escherichia coli rhamnose promoter
rhaP(BAD) is in Pseudomonas putida KT2440 independent of Crp-cAMP
activation. Appl Microbiol Biotechnol 85, 1923-1933.
430
431
432
433
Keller, K. L., Bender, K. S. & Wall, J. D. (2009). Development of a markerless
genetic exchange system for Desulfovibrio vulgaris Hildenborough and
its use in generating a strain with increased transformation efficiency.
Appl Environ Microbiol 75, 7682-7691.
434
435
436
437
Kobayashi, K., Ogura, M., Yamaguchi, H., Yoshida, K., Ogasawara, N., Tanaka,
T. & Fujita, Y. (2001). Comprehensive DNA microarray analysis of
Bacillus subtilis two-component regulatory systems. J Bacteriol 183,
7365-7370.
438
439
440
441
442
Krishnappa, L., Dreisbach, A., Otto, A., Goosens, V. J., Cranenburgh, R. M.,
Harwood, C. R., Becher, D. & van Dijl, J. M. (2013). Extracytoplasmic
proteases determining the cleavage and release of secreted proteins,
lipoproteins, and membrane proteins in Bacillus subtilis. J Proteome Res
12, 4101-4110.
443
444
445
Kristich, C. J., Manias, D. A. & Dunny, G. M. (2005). Development of a method
for markerless genetic exchange in Enterococcus faecalis and its use in
construction of a srtA mutant. Appl Environ Microbiol 71, 5837-5849.
446
447
448
449
Kunst, F., Ogasawara, N., Moszer, I., Albertini, A. M., Alloni, G., Azevedo, V.,
Bertero, M. G., Bessieres, P., Bolotin, A. & other authors (1997). The
complete genome sequence of the gram-positive bacterium Bacillus
subtilis. Nature 390, 249-256.
450
451
452
453
Leenhouts, K., Buist, G., Bolhuis, A., ten Berge, A., Kiel, J., Mierau, I.,
Dabrowska, M., Venema, G. & Kok, J. (1996). A general system for
generating unlabelled gene replacements in bacterial chromosomes. Mol
Gen Genet 253, 217-224.
454
455
Marahiel, M. A., Nakano, M. M. & Zuber, P. (1993). Regulation of peptide
antibiotic production in Bacillus. Mol Microbiol 7, 631-636.
456
457
458
459
Martínez-Garcia, E. & de Lorenzo, V. (2011). Engineering multiple genomic
deletions in Gram-negative bacteria: analysis of the multi-resistant
antibiotic profile of Pseudomonas putida KT2440. Environ Microbiol 13,
2702-2716.
460
461
462
Marx, C. J. & Lidstrom, M. E. (2002). Broad-host-range cre-lox system for
antibiotic marker recycling in gram-negative bacteria. Biotechniques 33,
1062-1067.
463
464
Michel, J. F. & Millet, J. (1970). Physiological studies on early-blocked
sporulation mutants of Bacillus subtilis. J Appl Bacteriol 33, 220-227.
465
466
467
468
Michna, R. H., Commichau, F. M., Tödter, D., Zschiedrich, C. P. & Stülke, J.
(2014). SubtiWiki - a database for the model organism Bacillus subtilis
that links pathway, interaction and expression information. Nucleic
Acids Res 42, D692-D698.
469
470
471
Molle, V., Fujita, M., Jensen, S. T., Eichenberger, P., Gonzalez-Pastor, J. E., Liu,
J. S. & Losick, R. (2003). The Spo0A regulon of Bacillus subtilis. Mol
Microbiol 50, 1683-1701.
472
473
474
475
Monteilhet, C., Perrin, A., Thierry, A., Colleaux, L. & Dujon, B. (1990).
Purification and characterization of the in vitro activity of I-Sce I, a novel
and highly specific endonuclease encoded by a group I intron. Nucleic
Acids Res 18, 1407-1413.
476
477
478
Mootz, H. D., Finking, R. & Marahiel, M. A. (2001). 4'-phosphopantetheine
transfer in primary and secondary metabolism of Bacillus subtilis. J Biol
Chem 276, 37289-37298.
479
480
481
482
Morimoto, T., Kadoya, R., Endo, K., Tohata, M., Sawada, K., Liu, S., Ozawa, T.,
Kodama, T., Kakeshita, H. & other authors (2008). Enhanced recombinant
protein productivity by genome reduction in Bacillus subtilis. DNA Res
15, 73-81.
483
484
485
486
Nakano, M. M., Corbell, N., Besson, J. & Zuber, P. (1992). Isolation and
characterization of sfp: a gene that functions in the production of the
lipopeptide biosurfactant, surfactin, in Bacillus subtilis. Mol Gen Genet
232, 313-321.
487
488
489
490
Paik, S. H., Chakicherla, A. & Hansen, J. N. (1998). Identification and
characterization of the structural and transporter genes for, and the
chemical and biological properties of, sublancin 168, a novel lantibiotic
produced by Bacillus subtilis 168. J Biol Chem 273, 23134-23142.
491
492
493
Petersohn, A., Brigulla, M., Haas, S., Hoheisel, J. D., Völker, U. & Hecker, M.
(2001). Global analysis of the general stress response of Bacillus
subtilis. J Bacteriol 183, 5617-5631.
494
495
496
Price, C. W., Fawcett, P., Cérémonie, H., Su, N., Murphy, C. K. & Youngman, P.
(2001). Genome-wide analysis of the general stress response in Bacillus
subtilis. Mol Microbiol 41, 757-774.
497
498
499
Reyrat, J. M., Pelicic, V., Gicquel, B. & Rappuoli, R. (1998). Counterselectable
markers: untapped tools for bacterial genetics and pathogenesis. Infect
Immun 66, 4011-4017.
500
501
502
Sambrook, J., Fritsch, E. F. & Maniatis, T. (1989). Molecular cloning : a
laboratory manual. Cold Spring Harbor, NY: Cold Spring Harbor
Laboratory.
503
504
Stein, T. (2005). Bacillus subtilis antibiotics: structures, syntheses and specific
functions. Mol Microbiol 56, 845-857.
505
506
507
508
Steller, S., Vollenbroich, D., Leenders, F., Stein, T., Conrad, B., Hofemeister, J.,
Jacques, P., Thonart, P. & Vater, J. (1999). Structural and functional
organization of the fengycin synthetase multienzyme system from
Bacillus subtilis b213 and A1/3. Chem Biol 6, 31-41.
509
510
Sun, T. & Altenbuchner, J. (2010). Characterization of a mannose utilization
system in Bacillus subtilis. J Bacteriol 192, 2128-2139.
511
512
Tamehiro, N., Okamoto-Hosoya, Y., Okamoto, S., Ubukata, M., Hamada, M.,
Naganawa, H. & Ochi, K. (2002). Bacilysocin, a novel phospholipid
513
514
antibiotic produced by Bacillus subtilis 168. Antimicrob Agents
Chemother 46, 315-320.
515
516
517
518
519
Tanaka, K., Henry, C. S., Zinner, J. F., Jolivet, E., Cohoon, M. P., Xia, F.,
Bidnenko, V., Ehrlich, S. D., Stevens, R. L. & other authors (2013).
Building the repertoire of dispensable chromosome regions in Bacillus
subtilis entails major refinement of cognate large-scale metabolic model.
Nucleic Acids Res 41, 687-699.
520
521
522
523
Tsuge, K., Ano, T., Hirai, M., Nakamura, Y. & Shoda, M. (1999). The genes degQ,
pps, and lpa-8 (sfp) are responsible for conversion of Bacillus subtilis
168 to plipastatin production. Antimicrob Agents Chemother 43, 21832192.
524
525
526
Turgay, K., Hahn, J., Burghoorn, J. & Dubnau, D. (1998). Competence in
Bacillus subtilis is controlled by regulated proteolysis of a transcription
factor. EMBO J 17, 6730-6738.
527
528
529
Wang, S. T., Setlow, B., Conlon, E. M., Lyon, J. L., Imamura, D., Sato, T., Setlow,
P., Losick, R. & Eichenberger, P. (2006). The forespore line of gene
expression in Bacillus subtilis. J Mol Biol 358, 16-37.
530
531
532
Warth, L. & Altenbuchner, J. (2013a). A new site-specific recombinasemediated system for targeted multiple genomic deletions employing
chimeric loxP and mrpS sites. Appl Microbiol Biotechnol 97, 6845-6856.
533
534
535
536
Warth, L. & Altenbuchner, J. (2013b). The tyrosine recombinase MrpA and its
target sequence: a mutational analysis of the recombination site mrpS
resulting in a new left element/right element (LE/RE) deletion system.
Arch Microbiol 195, 617-636.
537
538
539
Wenzel, M. & Altenbuchner, J. (2013). The Bacillus subtilis mannose regulator,
ManR, a DNA-binding protein regulated by HPr and its cognate PTS
transporter ManP. Mol Microbiol 88, 562-576.
540
541
542
543
Wenzel, M., Müller, A., Siemann-Herzberg, M. & Altenbuchner, J. (2011). Selfinducible Bacillus subtilis expression system for reliable and
inexpensive protein production by high-cell-density fermentation. Appl
Environ Microbiol 77, 6419-6425.
544
545
546
547
Westers, H., Dorenbos, R., van Dijl, J. M., Kabel, J., Flanagan, T., Devine, K. M.,
Jude, F., Seror, S. J., Beekman, A. C. & other authors (2003). Genome
engineering reveals large dispensable regions in Bacillus subtilis. Mol
Biol Evol 20, 2076-2090.
548
549
550
Westers, L., Westers, H. & Quax, W. J. (2004). Bacillus subtilis as cell factory
for pharmaceutical proteins: a biotechnological approach to optimize the
host organism. Biochim Biophys Acta 1694, 299-310.
551
552
553
Wood, H. E., Dawson, M. T., Devine, K. M. & McConnell, D. J. (1990).
Characterization of PBSX, a defective prophage of Bacillus subtilis. J
Bacteriol 172, 2667-2674.
554
555
556
Yanish-Perron, C., Vieira, J. & Messing, J. (1985). Improved M13 phage cloning
vectors and host strains: nucleotide sequences of the M13mp18 and
pUC19 vectors. Gene 33, 103-119.
557
558
559
560
561
Yoshida, K., Kobayashi, K., Miwa, Y., Kang, C. M., Matsunaga, M., Yamaguchi,
H., Tojo, S., Yamamoto, M., Nishi, R. & other authors (2001). Combined
transcriptome and proteome analysis as a powerful approach to study
genes under glucose repression in Bacillus subtilis. Nucleic Acids Res
29, 683-692.
562
563
564
Zahler, S. A., Korman, R. Z., Rosenthal, R. & Hemphill, H. E. (1977). Bacillus
subtilis bacteriophage SP: localization of the prophage attachment site,
and specialized transduction. J Bacteriol 129, 556-558.
565
566
567
568
569
Zakataeva, N. P., Nikitina, O. V., Gronskiy, S. V., Romanenkov, D. V. & Livshits,
V. A. (2010). A simple method to introduce marker-free genetic
modifications into the chromosome of naturally nontransformable
Bacillus amyloliquefaciens strains. Appl Microbiol Biotechnol 85, 12011209.
570
571
572
Zhang, X. Z., Yan, X., Cui, Z. L., Hong, Q. & Li, S. P. (2006). mazF, a novel
counter-selectable marker for unmarked chromosomal manipulation in
Bacillus subtilis. Nucleic Acids Res 34, e71.
573
574
575
576
577
578
Zheng, G., Hehn, R. & Zuber, P. (2000). Mutational analysis of the sbo-alb locus
of Bacillus subtilis: identification of genes required for subtilosin
production and immunity. J Bacteriol 182, 3266-3273.
579
Legends to figures
580
Figure 1: Plasmid map of the positive cloning and integration vector pJOE6743.1.
581
Selected double cutters for insertion of the amplified homologous regions were
582
added to the map. Binding sites for SP6 and T7 standard primers used for
583
sequencing are shown as white boxes. Abbreviations: manP (B. subtilis mannose
584
PTS transporter); ter (transcription terminator from phage ); amp (ampicillin
585
resistance); SpcR (spectinomycin resistance); lacPOZ: (lac promoter, lac operator
586
and lacZ  fragment).
587
Figure 2: Growth curves of B. subtilis 168 and IIG-Bs20-4 in LB and MG1 liquid
588
medium at 37°C (obtained in 4 independent experiments).
589
590
Table 1: Bacterial strains used in this study.
Strain
591
592
Genotype / Properties
Deletion from bp-bp*
Reference / Source
E. coli JM109
mcrA, recA1, supE44,
endA1, hsdR17, gyrA96,
relA1, thi, (lac-proAB),
F´[traD36, proAB+, lacIq
lacZM15]
n.a.
Yanisch-Peron et al. (1985)
B. subtilis 168
trpC2
n.a.
Bacillus Genetic Stock Center
B. subtilis 3NA
spo0A3
n.a.
Bacillus Genetic Stock Center
B. subtilis 
trpC2, SP, skin,
PBSX, pro1,
pks::CmR, pro3
n.a.
Westers et al. (2003)
IIG-Bs1
trp+
n.a.
this study
IIG-Bs2
manPA::erm
1272608..1275639
this study
IIG-Bs3
pro2
529495..604658
this study
IIG-Bs4
pro5
1879831..1886598
this study
IIG-Bs5
pro6
2055447..2079118
this study
IIG-Bs8
CmlR
1781877..1859956
this study
IIG-Bs9
subtilosin
3835926..3842999
this study
IIG-Bs10
plipastin
1960195..1997960
this study
IIG-Bs11
bacilysin
3867368..3874329
this study
IIG-Bs12
pro7
2708926..2742191
this study
IIG-Bs13
bacilysocin
3121673..3124005
this study
IIG-Bs14
Bacillus toxin
3464131..3467474
this study
IIG-Bs20
protease, sporulation
(sigE, sigG)
1599076..1606455
this study
IIG-Bs20-1
manPA::erm
1272643..1279510
this study
IIG-Bs20-2
pro4
1263599..1270416
this study
IIG-Bs20-3
3,3´-neotrehalosadiamine 1125000..1130795
this study
IIG-Bs20-4
srfAA-srfX (except comS) 377215..390803
391074..408169
this study
* n.a. (not applicable)
593
594
595
596
597
Table 2: Specific growth rates of B. subtilis 168 and IIG-Bs20-4 in LB and MG1
medium. The mean values with standard deviation is given obtained from four
independent experiments.
Strain
598
Maximum specific growth rate max [h-1]
LB
MG1
B. subtilis 168
1.42  0.14
0.63  0.09
IIG-Bs20-4
1.47  0.10
0.58  0.07