bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

- 1 The impact of CO2/HCO3 availability on anaplerotic flux in PDHC-

2 deficient Corynebacterium glutamicum strains

3

a a a a b# 4 Aileen Krüger , Johanna Wiechert , Cornelia Gätgens , Tino Polen , Regina Mahr

5 and Julia Frunzkea#

6

7 a Institut für Bio- und Geowissenschaften, IBG-1: Biotechnology, Forschungszentrum

8 Jülich, 52425 Jülich, Germany

9 b SenseUp GmbH, c/o Campus Forschungszentrum Jülich, 52425 Jülich, Germany

10

- 11 Running Head: Influence of CO2/HCO3 on anaplerotic flux

12

13 A.K. and J.W. contributed equally to this work.

14

15 #Address correspondence to Julia Frunzke, [email protected], and Regina

16 Mahr, [email protected].

17

1

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

18 Abstract

19 The complex (PDHC) catalyzes the oxidative

20 decarboxylation of pyruvate yielding acetyl-CoA and CO2. The PDHC-deficient

21 Corynebacterium glutamicum strain ΔaceE is therefore lacking an important

22 decarboxylation step in central . Additional inactivation of pyc, encoding

23 , resulted in a >15 hour lag phase in the presence of glucose,

24 while no growth defect was observed on gluconeogenetic substrates like acetate.

25 Growth was successfully restored by deletion of ptsG encoding the glucose-specific

26 permease of the PTS system, thereby linking the observed phenotype to the

27 increased sensitivity of strain ΔaceE Δpyc to glucose catabolism. In the following,

28 strain ΔaceE Δpyc was used to systematically study the impact of perturbations of

- 29 the intracellular CO2/HCO3 pool on growth and anaplerotic flux. Remarkably, all

- - 30 measures leading to enhanced CO2/HCO3 levels, such as external addition of HCO3

31 , increasing the pH, or rerouting metabolic flux via pentose phosphate pathway, at

32 least partially eliminated the lag phase of strain ΔaceE Δpyc on glucose medium. In

- 33 accordance, inactivation of the urease , lowering the intracellular CO2/HCO3

34 pool, led to an even longer lag phase accompanied with the excretion of L-valine and

35 L-alanine. Transcriptome analysis as well as an adaptive laboratory evolution

36 experiment of strain ΔaceE Δpyc revealed the reduction of glucose uptake as a key

37 adaptive measure to enhance growth on glucose/acetate mixtures. Altogether, our

- 38 results highlight the significant impact of the intracellular CO2/HCO3 pool on

39 metabolic flux distribution, which becomes especially evident in engineered strains

40 suffering from low endogenous CO2 production rates as exemplified by PDHC-

41 deficient strains.

42

2

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

43 Importance: CO2 is a ubiquitous product of cellular metabolism and an essential

44 substrate for carboxylation reactions. The pyruvate dehydrogenase complex (PDHC)

- 45 catalyzes a central metabolic reaction contributing to the intracellular CO2/HCO3 pool

46 in many organisms. In this study, we used a PDHC-deficient strain of

47 Corynebacterium glutamicum, which was additionally lacking pyruvate carboxylase

48 (ΔaceE Δpyc). This strain featured a >15 h lag phase during growth on glucose-

49 acetate mixtures. We used this strain to systematically assess the impact of

- 50 alterations in the intracellular CO2/HCO3 pool on growth on glucose-containing

- 51 medium. Remarkably, all measures enhancing the CO2/HCO3 levels successfully

- 52 restored growth emphasizing the strong impact of the intracellular CO2/HCO3 pool on

53 metabolic flux especially in strains suffering from low endogenous CO2 production

54 rates.

55

56

3

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

57 Introduction

58 CO2 is an inevitable product and at the same time an essential substrate of microbial

- 2- 59 metabolism. In water, CO2 is in equilibrium with HCO3 and CO3 , which is influenced

60 by the pH of the medium (1). As a product of decarboxylating reactions and substrate

- 61 for carboxylations, CO2 and bicarbonate (HCO3 ) are involved in various metabolic

- 62 processes. Consequently, the intracellular CO2/HCO3 pool has a significant impact

63 on metabolic fluxes like e.g. anaplerosis, especially in the early phase of cultivations

64 when cell density is low (2).

65 During aerobic growth, the pyruvate dehydrogenase complex (PDHC), which is

66 conserved in various microbial species, catalyzes an important reaction contributing

- 67 to the intracellular CO2/HCO3 pool (3, 4). The PDHC is a multienzyme complex

68 belonging to the family of 2-oxo acid dehydrogenase complexes also comprising the

69 2-ketoglutarate dehydrogenase complex (ODHC) and the branched-chain 2-ketoacid

70 dehydrogenase (3, 5). In particular, PDHC catalyzes the oxidative decarboxylation of

71 pyruvate to acetyl-CoA, thereby producing one molecule CO2 and one molecule

72 NADH per mol pyruvate.

73 The Gram-positive actinobacterium Corynebacterium glutamicum represents an

74 important platform strain used in industrial biotechnology for the production of amino

75 acids, proteins and various other value-added products (6-10). In this organism, the

76 PDHC complex represents a central target to engineer metabolic pathways of

77 pyruvate-derived products including L-valine, isobutanol, ketoisovalerate and L-

78 lysine. To this end, different studies focused on the reduction or complete

79 abolishment of PDHC activity to improve precursor availability (4, 11-13). Due to the

80 deficiency of PDHC activity, however, cells are not able to grow on glucose as single

81 carbon source which can be circumvented by the addition of acetyl-CoA refueling

82 substrates such as acetate. In presence of both carbon sources glucose and acetate, 4

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

83 PDHC-deficient strains initially form biomass from acetate and subsequently, convert

84 glucose into products (e.g. L-valine or L-alanine) in the stationary phase (12). In

85 contrast to Escherichia coli or Bacillus subtilis, C. glutamicum does not normally

86 show the typical diauxic growth behaviour, but prefers co-metabolization of many

87 carbon sources (10, 14). Co-utilization of glucose and acetate has been studied in

88 detail, showing that consumption rates of both carbon sources decrease, but the total

89 carbon consumption is comparable to growth on either carbon source alone which is

90 regulated by SugR activity (15). Further examples for co-metabolization of glucose

91 with other sugars or organic acids include fructose (16), lactate (17) as well as

92 pyruvate (14) and gluconate (18). It was shown, that upon growth on glucose-acetate

93 mixtures, the glyoxylate shunt is required to fuel the oxaloacetate pool in the wild type

94 (15). Here it is important to note that the glyoxylate shunt is active in the presence of

95 acetate, but repressed by glucose (19, 20).

96 For growth on glycolytic carbon sources, organisms depend on TCA cycle

97 replenishing reactions constituting the anaplerotic node (or phosphoenolpyruvate-

98 pyruvate-oxaloacetate node), comprised of different carboxylating and

99 decarboxylating reactions (21). Consequently, flux via these reactions is influenced

- 100 by the intracellular CO2/HCO3 pool. In contrast to most other organisms, C.

101 glutamicum possesses both anaplerotic carboxylases namely the

102 phosphoenolpyruvate carboxylase (PEPCx) (EC 4.1.1.31) encoded by the ppc gene

103 (22, 23) and the pyruvate carboxylase (PCx) (EC 6.4.1.1) encoded by pyc (21, 24).

104 These two C3-carboxylating catalyze bicarbonate-dependent reactions

105 yielding oxaloacetate from PEP or pyruvate, respectively. In C. glutamicum, it was

106 shown that these two enzymes could replace each other to a certain extent

107 depending on the intracellular concentrations of the respective effectors for each

108 enzyme. However, under standard conditions during growth on glucose, the biotin- 5

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

109 containing PCx contributes to the main anaplerotic activity of 90% compared to

110 PEPCx (24). Remarkably, the Michaelis-Menten constants of both carboxylases are

111 about 30-fold higher in comparison to Escherichia coli PEP carboxylase (25).

- 112 Apparently, a low CO2/HCO3 pool may thus limit anaplerotic flux, which is of special

113 relevance in early phases when biomass concentration is low, but high aeration may

114 strip dissolved CO2 from the medium.

115 Several studies revealed the inhibitory effect of low pCO2 on microbial growth (2, 26,

116 27). In the case of C. glutamicum, low pCO2 pressures led to a significant drop in

117 growth rate in turbidostatic continuous cultures (28) as well as batch cultures (29). It

118 was further demonstrated that the reduced flux through under

- 119 low CO2/HCO3 levels led to an increased production of the pyruvate-derived amino

120 acids L-alanine and L-valine (29).

121 While several previous studies focused on the impact of CO2 by altering the CO2

122 partial pressure in the process, we applied targeted genetic perturbations to

- 123 systematically assess the impact of the intracellular CO2/HCO3 pool on anaplerotic

124 flux and growth of C. glutamicum. Here, we focused on PDHC-deficient strains

- 125 suffering from a reduced intracellular CO2/HCO3 pool during growth on glucose-

- 126 acetate mixtures. The effect of low CO2/HCO3 levels became even more evident,

127 upon additional deletion of the pyc gene encoding the dominant anaplerotic enzyme

128 leading to a drastically elongated lag phase of a C. glutamicum ΔaceE Δpyc strain

129 during growth on glucose acetate mixtures. This effect could be attributed to a

130 reduced tolerance of these strains towards glucose and was successfully

131 complemented by deletion of ptsG encoding the glucose-specific permease of the

132 PTS system. Remarkably, growth was successfully restored by increasing the

- 133 intracellular CO2/HCO3 pool by the addition of bicarbonate, by increasing the pH, by

134 re-routing metabolic flux over the pentose phosphate pathway or by refueling of the 6

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

135 oxaloacetate pool by the addition of TCA intermediates to the growth medium.

136 Finally, an adaptive laboratory evolution (ALE) of C. glutamicum ΔaceE Δpyc strain

137 on minimal medium containing glucose and acetate revealed the abolishment of

138 glucose uptake as a primary strategy to allow for growth on acetate.

139

140

7

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

141 Results

- 142 A lowered intracellular CO2/HCO3 pool impacts growth of PDHC-deficient

143 strains

144 The pyruvate dehydrogenase complex (PDHC) comprises a key metabolic reaction in

145 the central metabolism by catalyzing the oxidative decarboxylation of pyruvate to

146 acetyl-CoA (Fig. 1), thereby producing one molecule CO2 and one molecule NADH

147 per mol pyruvate. Previous studies already revealed that PDHC-deficient strains

148 feature an about two-fold decreased excretion of CO2 during exponential growth (30).

- 149 It is therefore likely to assume that a reduction of the intracellular CO2/HCO3 pool

150 has an impact on metabolic flux distribution in PDHC-deficient strains. A simple

151 growth comparison of the C. glutamicum wild type and a strain lacking the ΔaceE

152 gene (thus PDHC-deficient) revealed a slightly increased lag phase of the ΔaceE

153 strain which especially became evident when cells were grown in microtiter plates,

154 while the maximal growth rate appeared to be unaffected (ΔaceE: 0.40 ± 0.01 h-1;

155 WT: 0.41 ± 0.01 h-1) (Fig. 2A). Remarkably, this lag phase was completely eliminated

- 156 by the addition of 100 mM HCO3 attributing the effect to the lowered intracellular

- 157 CO2/HCO3 pool of this mutant (Fig. 2B).

158 In C. glutamicum, the pyruvate carboxylase (PCx) encoded by pyc is the dominating

- 159 HCO3 -dependent anaplerotic enzyme required for refueling the TCA cycle via

160 oxaloacetate during growth on glycolytic carbon sources (24). In the following, we

161 deleted pyc in the ΔaceE background to examine the influence of different

- 162 CO2/HCO3 levels on bicarbonate-dependent anaplerosis. Remarkably, the additional

163 deletion of the pyc gene resulted in an elongated lag phase of about 20 hours and a

-1 164 reduced growth rate of strain ΔaceE Δpyc (µmax = 0.27 ± 0.02 h compared to 0.40 ±

165 0.01 h-1 of strain ΔaceE) during cultivation in CGXII minimal medium containing

166 glucose and acetate (Fig. 2A). An extended lag phase was also observed for the 8

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

167 Δpyc single mutant, but the lag phase was significantly less prominent compared to

168 the PDHC-deficient variant ΔaceE Δpyc (Fig. 2C). The growth defect of both strains

169 was successfully complemented by re-introducing the pyc gene on the plasmid pJC1

170 under control of its own promoters (Fig. 2 C). Remarkably, this extended lag phase

171 was only observed in the presence of glucose, but not during cultivation on minimal

172 medium containing acetate as sole carbon source (Fig. 2A).

173

174 Glucose uptake results in strongly retarded growth in strains lacking pyruvate

175 carboxylase

176 In order to verify whether glucose uptake caused the retarded growth, the ΔaceE

177 Δpyc strain was cultivated in the presence of 154 mM acetate and increasing

178 amounts of glucose (100-250 mM). In line with our hypothesis, the lag phase showed

179 a stepwise increase when higher amounts of glucose were added to the medium

180 (Fig. 3A). Consistently, ΔaceE Δpyc also showed significant elongated lag phases on

181 the PTS sugars fructose and sucrose (Fig. S1). By contrast, the glucose

182 concentration did not significantly affect the growth of the ΔaceE strain (Fig. 3A), in

183 which the major anaplerotic enzyme PCx still replenishes the oxaloacetate pool.

184 To link the observed growth phenotype to the uptake of glucose, we deleted the ptsG

185 gene, encoding the glucose specific permease of the PTS system, in the ΔaceE and

186 ΔaceE Δpyc strain background. Growth and glucose consumption rates of resulting

187 strains were analyzed during growth on glucose and acetate. Remarkably, deletion of

188 ptsG fully restored the growth of the parental strain ΔaceE Δpyc (ΔaceE Δpyc: 0.27 ±

189 0.02 h-1; ΔaceE Δpyc ΔptsG: 0.39 ± 0.02 h-1), but resulted in reduced final

190 backscatter values comparable to growth on acetate alone (Fig. 3B, Table 3). Strains

191 lacking the ptsG gene consumed only minor amounts of glucose, while strains ΔaceE

192 and ΔaceE Δpyc consumed glucose after entering the stationary phase (Fig. S2). 9

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

193 Glucose uptake rates of ΔaceE Δpyc were significantly reduced to 10.87 ± 4.95 nmol

194 min-1 g-1 compared to ΔaceE with 16.81 ± 3.01 nmol min-1 g-1. In contrast, deletion of

195 ptsG did not restore growth on the PTS substrate fructose (Fig. S3), which is mainly

196 imported by the permease encoded by ptsF (31, 32). Altogether, these findings

197 clearly link the observed growth defect of the ΔaceE Δpyc strain to the uptake of

198 glucose or alternative PTS substrates.

199

- 200 Increased HCO3 availability improved anaplerotic flux in pyc mutants

- 201 In the following, we used different strategies to alter the intracellular CO2/HCO3 pool.

- 202 This was achieved by (1) supplementing the medium with HCO3 , (2) by mutation of

203 the endogenous urease gene producing CO2 from urea in the early growth phase, (3)

204 by increasing the pH of the growth medium, and (4) by re-routing metabolic flux via

205 the pentose phosphate pathway (Fig. 3).

- 206 Similar as observed for the parental strain ΔaceE, addition of 100 mM HCO3

207 eliminated the lag phase of PDHC-deficient strains lacking the pyc gene (Fig. 2B)

208 during microtiter plate cultivation on glucose-acetate mixtures. Restoration of growth

209 was not possible in a strain lacking both anaplerotic enzymes PCx and PEPCx,

- 210 indicating that the positive effect of HCO3 on the ΔaceE Δpyc strain depends on the

211 increased flux over PEPCx (Fig. 4A, Fig. S4). In contrast to PCx-deficient strains, no

212 negative impact of glucose on growth of the strain ΔaceE Δppc was observed (Fig.

213 S5) confirming again the superior role of PCx in C. glutamicum (24, 33). Interestingly,

214 we also observed a significant impact of the culture volume on growth of ΔaceE

215 Δpyc. Since efficient liquid-air mixing will continuously remove CO2 from the culture

216 medium, the lag phase of ΔaceE Δpyc was much more prominent during microtiter

217 plate cultivation (~17 h) compared to cultures in shake flasks (~7 h lag phase in 100

218 ml cultures; ~13 h lag phage in 50 ml) (Fig. S6). 10

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

219 The urease enzyme, catalyzing the degradation of urea to ammonium and carbon

- 220 dioxide, represents a further important contributor to the intracellular CO2/HCO3 pool,

221 especially in the early exponential phase when cell densities and decarboxylation

222 reaction rates are low. In a previous study, the ureD-E188* mutation was found to

223 abolish urease activity in C. glutamicum and to enhance L-valine production of

224 PDHC-deficient strains (34). In this study, we introduced the ureD-E188* mutation

225 into the ΔaceE Δpyc background to examine the influence of an even lower

- 226 CO2/HCO3 level on anaplerosis in this particular strain background. The additional

227 inactivation of the urease enzyme resulted in a significantly elongated lag phase of

- 228 about 46 hours, which could also be complemented by addition of HCO3 (Fig. 4B)

- 229 demonstrating the importance of the CO2/HCO3 pool during the initial growth phase.

230 Furthermore, we assessed the impact of urease mutation on L-valine and L-alanine

231 production of strain ΔaceE and strain ΔaceE ureD-E188*and the complementation

- 232 via addition of HCO3 . Within 28 hours of cultivation, the ΔaceE strain produced 15

233 mM L-valine and 10 mM L-alanine as by-product, while the ΔaceE ureD-E188*

- 234 accumulated 54 mM L-valine and 38 mM L-alanine (Fig. 5). Addition of HCO3 ,

235 however, significantly reduced L-valine production by 54% (ΔaceE) and 59% (ΔaceE

236 ureD-E188*), respectively. The by-product L-alanine increased by 22% for the ΔaceE

237 strain and decreased by 51% for the ΔaceE ureD-E188* strain (Fig. 5). The example

238 of L-valine/L-alanine production illustrates the important role of the intracellular

- 239 CO2/HCO3 levels on metabolic flux in PDHC-deficient strains.

- 240 Another approach to influence the extracellular HCO3 availability was to change the

241 pH conditions (2). According to the Bjerrum plot, CO2 is predominant at acidic

2- 242 conditions, while at alkaline conditions CO3 is the dominant form. At a pH of about

- - 243 8, HCO3 is the prevalent form. As HCO3 is the substrate of the anaplerotic enzymes

244 PCx and PEPCx, the pH in the cultivation medium was adjusted to pH 8 using KOH 11

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

- 245 leading to an equilibrium shift towards higher HCO3 availability. As negative control,

246 the pH was adjusted to pH 6 using HCl. While growth of both strains was retarded at

247 pH 6, it was significantly enhanced at pH 8 (Fig. 4C). Thus, elevation of the culture

248 pH improved growth of PDHC-deficient strains likely by increasing anaplerotic flux.

249 With the reaction catalyzed by the 6-phosphogluconate dehydrogenase (6PGDH),

250 the PPP comprises another decarboxylation reaction in central carbon metabolism,

- 251 thereby contributing to the intracellular CO2/HCO3 pool. Remarkably, the lag phases

252 of ΔaceE Δpyc and ΔaceE ureD-E188* Δpyc were mostly complemented during

253 growth on acetate and gluconate – the latter entering the PPP via gluconate-6-

254 phosphate (Fig. 4D). Growth on ribose, which enters the PPP via ribulose-5-

255 phosphate and thereby bypassing the decarboxylation catalyzed by 6PGDH, showed

256 a significantly elongated lag phase in pyc deficient strains (Fig. 4D). In a further

257 attempt, metabolic/glycolytic flux was re-routed through the PPP by deleting pgi,

258 encoding the glucose-6-phosphate isomerase. However, deletion of pgi resulted in a

259 severe growth defect in the ΔaceE background. In this context the growth of ΔaceE

260 Δpyc featured only a minor, non-significant improvement (Fig. S7).

261 Taken together, our findings highlight the important impact of the intracellular

- 262 CO2/HCO3 pool on metabolic flux in central carbon metabolism, which is especially

263 evident in strain ΔaceE Δpyc lacking a central decarboxylation reaction and the key

264 carboxylase PCx in C. glutamicum.

265

266 Refueling the TCA cycle improves growth of pyc mutants

267 Glucose catabolism requires sufficient anaplerotic flux to replenish TCA cycle

268 intermediates providing precursors for various anabolic pathways. Therefore, we

269 tested whether the addition of TCA intermediates would complement the negative

270 impact of glucose on the growth of ΔaceE Δpyc. Remarkably, all tested TCA 12

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

271 intermediates (succinate, malate, citrate and the glutamate containing dipeptide Glu-

272 Ala) reduced the extended lag phase of ΔaceE Δpyc during growth on glucose-

273 acetate mixtures (Fig. 6A). The dipeptide Glu-Ala and succinate reduced the

274 elongated lag phase to a high extent by 90 % (Δt = 10 %) and 85 % (Δt = 15 %),

275 respectively, while the effects of citrate and malate were weaker (Δt = 40 % and 38

276 %, respectively) (Table 4).

277 Based on previous studies, it was known that the glyoxylate shunt might be switched

278 off during growth on glucose-acetate mixtures, due to the inhibitory effect of glucose

279 (35). Here, overexpression of the genes aceA and aceB, encoding the glyoxylate

280 shunt enzymes isocitrate lyase and malate synthase, resulted in a significantly

281 shorter lag phase of strain ΔaceE Δpyc (Fig. 6B). Overexpression was accomplished

282 using the vector pJC1 harboring an inducible Ptac promoter in front of two synthetic

283 operon variants: pJC1-aceA-linker-aceB and pJC1-aceB-linker-aceA. However, it has

284 to be noted that the overexpression of aceA and aceB led to slightly longer lag

285 phases of strain ΔaceE than compared to the empty vector control (Fig. 6B).

286 Expression from the leaky Ptac promoter yielded the best result, while induction with

287 IPTG resulted in severe growth defects (data not shown).

288 Unusual intermediate accumulation or depletion can provide valuable information

289 regarding intracellular flux imbalances causing the observed growth defects. Gas

290 chromatography-time of flight (GC-ToF) analysis was performed for analysis of the

291 metabolic states by comparing samples of ΔaceE during early exponential phase,

292 ΔaceE Δpyc at the same time point during the lag phase, as well as ΔaceE Δpyc

293 during early exponential phase (see Fig. S8 showing the sampling scheme).

- 294 Additionally, samples of both PDHC-deficient strains cultivated with 100 mM HCO3

295 were measured and WT cells during exponential phase were used as reference

296 (Table S3). No significant intermediate accumulation or occurrence of unusual 13

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

297 compounds was observed in the lag phase sample of ΔaceE Δpyc compared to all

298 other samples. Although oxaloacetate cannot be measured by this method, a

299 complete depletion of oxaloacetate-derived aspartate was detected in ΔaceE Δpyc

- 300 cells during the lag phase but not during exponential phase or in HCO3

301 complemented samples. This supports the hypothesis that oxaloacetate depletion is

302 the key reason for the elongated lag phase.

303 In order to identify oxaloacetate depletion as key growth limiting factor and to

304 decouple this effect from anaplerosis-dependent replenishment, strains Δpyc, Δpyc

305 Δppc and ΔaceE Δpyc Δppc were tested in the presence of different TCA

306 intermediates. Here, mutants lacking both anaplerotic reactions (PCx and PEPCx,

307 strains Δpyc Δppc and ΔaceE Δpyc Δppc) were not able to grow on glucose acetate

308 mixtures at all, but grew on gluconeogenetic substrates like acetate (Fig. 4A, Fig.

309 S4). Again, all TCA intermediates were able to complement the glucose sensitivity of

310 these pyc-deficient strains and the growth of Δpyc Δppc and ΔaceE Δpyc Δppc was

311 effectively restored (Table 5). Nevertheless, slight differences were observed, which

312 might also be caused by differences in the uptake of these carbon sources.

313

314 Adaptation of C. glutamicum ΔaceE Δpyc to growth on glucose and acetate

315 In a previous study, Kotte et. al reported on an elongated lag phase of E. coli as a

316 result of glucose-gluconeogenetic substrate shifts (36). This phenomenon was

317 ascribed to the formation of a small subpopulation, which is able to start growing after

318 carbon source switches (36). Usually, the growth of a small subpopulation is

319 obscured by typical bulk measurements, but can be visualized by single-cell

320 approaches, such as flow cytometry. To this end, the membranes of ΔaceE and

321 ΔaceE Δpyc cells were stained using the non-toxic, green fluorescent dye PKH67

322 prior to inoculation. By cellular growth, the amount of fluorescent dye is diluted by 14

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

323 membrane synthesis leading to a decrease of single cell PKH67 fluorescence (36).

324 Stained ΔaceE and ΔaceE Δpyc cells were cultivated in minimal medium containing

325 glucose and acetate (Fig. 7A). As non-proliferating control, ΔaceE Δpyc cells were

326 incubated in glucose as sole carbon source. Using flow cytometry, membrane

327 staining was analyzed during the cultivation at the single-cell level (Fig. 7B). While

328 the mean of fluorescence of the whole ΔaceE population decreased from 2.7 × 104

329 a.u. to 7.5 × 102 a.u., the mean of the ΔaceE Δpyc population incubated in glucose

330 medium only shifted from 3.1 × 104 a.u. to 1.3 × 103 a.u. resulting from a minor

331 dilution/bleaching of fluorescence intensity during incubation. Remarkably, both

332 strains, ΔaceE and ΔaceE Δpyc, featured a rather heterogeneous adaptation

333 behavior on glucose and acetate mixtures which was apparently time delayed in

334 strain ΔaceE Δpyc reflecting the elongated lag phase. Already in the early phase of

335 cultivation, the populations split up in two subpopulations of low and high PKH67

336 fluorescence indicating that only a few cells start to proliferate, while a fraction of the

337 population was not able to grow under same conditions (37-39).

338 To study the adaptation of ΔaceE Δpyc, comparative transcriptome analysis were

339 performed using DNA microarrays. To this end, ΔaceE and ΔaceE Δpyc cells were

340 harvested at a comparable optical density during early exponential phase (Fig. S8).

341 Both strains were cultivated in CGXII media with 154 mM acetate and 222 mM

342 glucose. Under the chosen conditions, a total of 354 genes showed a significantly

343 altered mRNA level. While 121 genes were at least 1.7-fold upregulated (p-value <

344 0.05), 233 genes were 0.7-fold downregulated (p-value < 0.05) in strain ΔaceE Δpyc

345 (Table S4). An overview of expression changes of selected genes is shown in Table

346 6. Among the various changes on the level of gene expression, we observed a 0.59-

347 fold and 0.67-fold downregulation of the respective PTS-system components ptsI and

348 ptsG in strain ΔaceE Δpyc, which is in line with the decreased glucose uptake rates 15

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

349 of this strain (Table 3). In contrast, iolT1 encoding the myo-inositol transporter, which

350 is inter alia responsible for PTS-independent glucose uptake, was upregulated (40,

351 41). This is in line with the fact that the ptsG deficient strains still consumed minor

352 amounts of glucose. Interestingly, genes involved in acetate metabolism, including

353 aceA and aceB constituting the glyoxylate shunt, displayed an increased mRNA level

354 in strain ΔaceE Δpyc. This may also represent an important adaptive mechanism, as

355 the glyoxylate shunt needs to be active in ΔaceE Δpyc to compensate for the loss of

356 TCA intermediates during co-consumption of acetate and glucose as demonstrated

357 by the effect of aceA-aceB overexpression in this study (Fig. 6B). While the mRNA

358 level of genes encoding anaplerotic enzymes did not show a significant change, the

359 glycolytic genes pgi and pyk featured an about two-fold reduced level in the ΔaceE

360 Δpyc mutant. This might as well reflect an adaptive mechanism, as a downregulation

361 of pyk would reduce flux from PEP to pyruvate and reduced expression of pgi would

362 probably lead to an increased flux through the PPP contributing an additional

363 decarboxylation step.

364

365 Less is more - Adaptive laboratory evolution of ΔaceE Δpyc revealed rapid

366 inactivation of glucose uptake

367 In order to identify key mutations abolishing the lag phase of ΔaceE Δpyc, an

368 adaptive laboratory evolution (ALE) experiment was performed. In this ALE,

369 C. glutamicum ΔaceE Δpyc was grown in CGXII minimal medium containing 154 mM

370 acetate and 222 mM glucose in overall 16 repetitive-batch cultures. Remarkably,

371 already after eight inoculations a significantly shortened lag phase was observed;

372 after 10 inoculations strain ΔaceE Δpyc featured a similar lag phase as the strain

373 ΔaceE (Fig. 8A). In contrast, serial cultivation transfers in CGXII medium containing

374 solely acetate did not lead to improved growth in glucose-acetate medium indicating 16

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

375 that glucose acted as evolutionary selection pressure (Fig. S9). Genome re-

376 sequencing of three selected glucose-acetate clones presented in Fig. 8B revealed

377 mutations in the ptsI gene, encoding the EI enzyme of the PTS system, in two

378 independently evolved clones (Table 7). This is in agreement with the reduced

379 biomass formation of both clones, which apparently abolished glucose uptake to

380 optimize growth on acetate. The third clone showed only a slightly accelerated

381 growth, but was not impaired in biomass formation. Here, sequencing revealed a

382 mutation in the rpsC gene, encoding the ribosomal S3 protein.

383 In summary, it was indeed possible to eliminate the observed lag phase of strain

384 ΔaceE Δpyc ascribed to glucose sensitivity by an adaptive laboratory evolution

385 approach. Especially, the mutations identified in the ptsI gene are in agreement with

386 other experiments emphasizing the high selective pressure exerted by glucose on

- 387 PCx-deficient strains when the intracellular HCO3 levels are limiting.

388

389

17

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

390 Discussion

391 During growth on glucose, the pyruvate dehydrogenase complex catalyzes a central

- 392 metabolic reaction contributing to the intracellular CO2/HCO3 pool. In this study, we

- 393 systematically perturbed the intracellular CO2/HCO3 pool of PDHC-deficient

394 C. glutamicum strains to monitor the impact on growth and anaplerotic flux. Even the

395 PDHC-deficient strain C. glutamicum ΔaceE features a slightly elongated lag phase

396 on glucose-acetate mixtures compared to the wild type, while no difference in growth

397 is observed when strains are growing on gluconeogenetic substrates, like acetate.

398 This is remarkable since this strain catabolizes only a minor fraction of the provided

399 glucose in the exponential growth phase, but rather turns to the production of L-

400 valine and L-alanine from glucose in the stationary phase (11, 12). The growth defect

401 of C. glutamicum ΔaceE was, however, successfully complemented by the addition of

- 402 HCO3 which had no significant effect on growth of wild type cells, therefore, hinting

403 towards problems caused by an impaired anaplerotic flux in this strain background.

404 By comparing the CO2 production rates of the C. glutamicum ΔaceE strain and the

405 wild type, Bartek et al. could show in a previous study, that strain ΔaceE excretes

-1 -1 -1 406 only around 0.83 mmol g h of CO2, while the wild type excretes ca. 1.65 mmol g

-1 407 h CO2 (30). The lower CO2 production rates of the PDHC-deficient strain already

- 408 indicate a significant impact on the intracellular CO2/HCO3 pool affecting metabolic

- 409 flux distribution especially at low cell densities when CO2/HCO3 is limiting.

410 The anaplerotic node comprises the essential link between

411 glycolysis/gluconeogenesis and the TCA cycle (21). In contrast to most other

412 organisms, C. glutamicum possesses both anaplerotic carboxylases, PEPCx

413 (encoded by ppc) and PCx (encoded by pyc) catalyzing the anaplerotic bicarbonate-

414 dependent reactions to yield oxaloacetate from PEP or pyruvate, respectively. In C.

415 glutamicum, it was shown that these two enzymes can replace each other to a 18

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

416 certain extent depending on the intracellular concentrations of the respective

417 effectors for each enzyme. However, during growth on glucose, the biotin-containing

418 PCx contributes to the main anaplerotic activity of 90% compared to PEPCx (24, 33).

419 In our study, deletion of the pyc gene resulted in a severely elongated lag phase (>15

420 hours) of ΔaceE Δpyc strains when glucose was present in the medium. Retarded

421 growth of Δpyc strains have been reported previously and, in line with our data,

422 Blombach and co-workers also reported a severe growth defect of this strain under

- 423 low CO2/HCO3 levels (24, 29). This effect appeared to be even more severe when

424 the cells were grown in microtiter plates under CO2-stripping conditions (low culture

425 volume, high air mixing) compared to growth in higher culture volumes (Fig. S6). The

426 PDHC-deficient mutant strain C. glutamicum ΔaceE Δpyc was found to react very

427 sensitively towards small changes in the intracellular bicarbonate availability and was

428 therefore used as a test platform to systematically assess the impact of perturbations

- - 429 affecting the intracellular CO2/HCO3 pool. While the addition of HCO3 , the increase

430 of the pH as well as higher culture volumes rescued the strain, mutation of the urease

- 431 accessory protein (ureD-E188*), which lowers the intracellular CO2/HCO3 pool,

432 resulted in an even more severe phenotype. It was not possible to restore growth if

433 the second anaplerotic gene ppc was deleted as well, confirming that the

434 abovementioned measures fostered the flux over the remaining anaplerotic reaction

435 catalyzed by PEPCx.

436 The mutation in ureD was revealed by a previous biosensor-driven evolution

437 approach selecting for mutations enhancing L-valine production in C. glutamicum

438 (34). Here, inactivation of the urease enzyme by the mutation ureD-E188* reduced

439 the anaplerotic flux via PCx resulting in an increased precursor availability of

440 pyruvate-derived products like L-valine and L-alanine (34). In the present study, we

- 441 further confirmed this finding by complementation with HCO3 counteracting the effect 19

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

- 442 of the ureD-E188* mutation. In another study by Blombach et al., lower CO2/HCO3

443 levels also triggered enhanced production of L-valine and L-alanine of C. glutamicum

444 during shake flask cultivation (29). However, a positive impact of anaplerotic

445 reactions has been confirmed for other, for example, oxaloacetate-derived products

446 like L-lysine. Here, attempts to overexpress pyc or the introduction of deregulated

447 variants significantly improved L-lysine production (42, 43). However, in spite of great

448 efforts and successes in the development of L-lysine strains, the impact of altered

- 449 CO2/HCO3 levels has not been systematically assessed so far.

450 Residual glucose consumption of PDHC-deficient C. glutamicum strains has already

451 been reported in previous studies (11, 12, 30). The results of this work, however,

452 emphasize that mutants lacking the major anaplerotic route via PCx are under strong

453 evolutionary pressure in the presence of glucose. Although strains would be capable

454 to grow on the acetate supplied in the growth medium, increasing levels of glucose

455 resulted in a severely elongated lag phase (up to >40 hours with 250 mM) of C.

456 glutamicum ΔaceE Δpyc. Besides the abovementioned efforts to increase the

- 457 CO2/HCO3 pool, also the deletion of the ptsG gene itself effectively restored growth

458 on glucose, but not on other PTS substrates. This indicates that this strain is

459 sensitive to glucose, but does not have issues with regulatory functions of PtsG

460 known from e.g. E. coli (44, 45).

461 Furthermore, among the three key mutations identified in the ALE experiment, two

462 SNPs obtained from independent cell lines were located within the ptsI gene,

463 encoding the EI component of the PTS system. These findings clearly highlight the

464 problems of residual glucose consumption in strains featuring an impaired anaplerotic

465 flux. One reason for the observed growth phenotype might be stress resulting from

466 the intracellular accumulation of sugar phosphates. In Escherichia coli, sugar

467 phosphate stress can be caused by accumulation of any sugar phosphates due to a 20

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

468 block in glycolysis (46), e.g. by mutations in pgi or pfkA (47), or due to excessive

469 glucose uptake caused e.g. by overexpression of uhpT encoding a sugar phosphate

470 permease (48). The resulting metabolic imbalance causes growth inhibition. For

471 example, in C. glutamicum, accumulation of several sugar phosphates, like fructose-

472 1,6-bisphosphate or PEP, inhibits the isocitrate lyase (35). To counteract this stress,

473 E. coli triggers SgrR which consequently activates SgrS by an unknown signal. The

474 transcription factor SgrS then prevents further uptake of glucose by downregulation

475 of ptsG (46, 49, 50). This is in line with the finding of our study, showing a slight

476 downregulation of ptsG upon resumed growth of strain ΔaceE Δpyc (Table 3). This

477 reduction of glucose uptake via the PTS system which converts PEP to pyruvate

478 during glucose transport also contributes to an increase in PEP availability for the

479 remaining anaplerotic reaction catalyzed by PEPCx. Further, this is in line with an

480 upregulation of iolT1, which ensures minor usage of glucose without PEP depletion

481 (40, 41). In this context, also the downregulation of the pyk gene in strain ΔaceE

482 Δpyc can be interpreted as a potential adaptation to increase the PEP pool fostering

483 anaplerotic flux.

484 During growth on gluconeogenic carbon sources - like acetate - the transcriptional

485 repressor SugR represses the expression of the PTS genes in C. glutamicum,

486 including the glucose-specific ptsG gene, but also ptsF, ptsS and general

487 components as ptsI and ptsH (51-53). Derepression appears to be triggered by the

488 accumulation of fructose-6-phosphate (F6P), which is generated from glucose 6-

489 phosphate entering glycolysis. Consequently, F6P accumulation leads to an increase

490 of glucose consumption rates resulting in parallel catabolization of glucose and

491 acetate (15). The fact that the PDHC-deficient strains are also subject to regulation of

492 SugR was demonstrated by studies showing that ΔaceE ΔsugR strain leads to up to

493 4-fold higher glucose consumption rates (54). 21

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

494 In aerobic glucose-based bioprocesses the endogenous production of CO2 is

495 typically sufficiently high to promote microbial growth even at low cell densities.

496 However, it is a well-known fact that anaerobic growth of some bacteria, like E. coli,

- 497 requires the external supply of CO2/HCO3 to avoid long lag phases (26, 55, 56). The

- 498 critical impact of CO2/HCO3 levels may become especially evident under conditions

499 when the endogenous CO2 production rate becomes limiting. This was nicely

500 demonstrated by a recent study of Bracher et al. who could show that long lag

501 phases of engineered, but non-evolved Saccaromyces cerevisiae strains during

502 xylose fermentation could be avoided by sparging the bioreactor cultures with CO2/N2

503 mixtures (57). Alternatively, the addition of L-aspartate, whose transamination

504 provides oxaloacetate refueling the TCA, completely abolished the long adaptation

505 phase of the respective yeast strains. In line with these findings, a recent 13C-flux

506 analysis with E. coli revealed that a considerable turnover of lipids via β-oxidation

- 507 appears to be required for growth on xylose to enhance the intracellular CO2/HCO3

508 pool to a growth-promoting level enabling anaplerotic flux (58). These findings are in

509 nice agreement with our study showing that the elongated lag phase of strain C.

510 glutamicum ΔaceE Δpyc can be eliminated by different measures enhancing either

- 511 the intracellular CO2/HCO3 level or by refueling of the TCA cycle with different

512 intermediates, like citrate, malate or succinate. Overall, oxaloacetate depletion

513 appeared to represent a key issue causing the delayed growth of the ΔaceE Δpyc

514 strain. Our data revealed that this is caused by an impaired anaplerotic flux on

- 515 glucose due to low intracellular CO2/HCO3 levels as well as reduced activities of

516 glyoxylate enzymes.

517 Altogether, these results emphasize the important impact of the intracellular

- 518 CO2/HCO3 pool for metabolic flux distribution, which is especially relevant in

519 engineered strains suffering from lower endogenous CO2 production rates as 22

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

520 exemplified by PDHC-deficient strains in this study, but also by the performance of

521 pentose-fermenting yeast and E. coli strains (57, 58). Consequently, the lack of an

522 important by-product, like CO2 released by the PDHC complex, may have a

523 significant impact on cellular metabolism and growth, especially on glycolytic

524 substrates demanding a high flux via the anaplerotic reactions.

525

23

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

526 Material and methods

527 Bacterial strains and growth conditions

528 All bacterial strains and plasmids used in this study are listed in Table 1. Mutant

529 strains are based on the Corynebacterium glutamicum ATCC 13032 wild type strain

530 (59).

531 Standard cultivation of C. glutamicum ΔaceE cells and derivatives were performed on

532 BHI (brain heart infusion, Difco BHI, BD, Heidelberg, Germany) agar plates

533 containing 51 mM potassium acetate (hereinafter named acetate) (ChemSolute, Th.

534 Geyer, Stuttgart, Germany) at 30°C for two days. One single colony was picked and

535 incubated for 8 to 10 h at 30°C in either 4.5 ml or 1 ml BHI containing 154mM acetate

536 in reaction tubes or Deepwell-plates (VWR International, Pennsylvania, USA),

537 respectively. First pre-cultures were used to inoculate second pre-cultures in CGXII

538 minimal medium (60) supplemented with 154 mM acetate either as 10 ml culture in

539 shake flasks or in 1 ml volume in Deepwell-plates. After overnight growth, a main

540 culture was inoculated at an OD600 of 1 in CGXII media containing 154 mM acetate

541 and either 222 mM D(+)-glucose monohydrate (Riedel-de Haën, Seelze, Germany)

542 (further referred to as glucose), or any other carbon source as stated, e.g. D(-)-

543 fructose (Sigma-Aldrich, Taufkirchen, Germany) (referred to as fructose), D(+)-

544 sucrose (Roth, Karlsruhe, Germany) (referred to as sucrose), D-gluconic acid sodium

545 salt (Sigma-Aldrich, Taufkirchen, Germany) (referred to as gluconate), D(-)ribose

546 (Sigma-Aldrich, Taufkirchen, Germany) (referred to as ribose), citric acid

547 monohydrate (Merck Millipore, Darmstadt, Germany) (referred to as citrate), H-Glu-

548 Ala-OH (Bachem AG, Bubendorf, Germany) (referred to as Glu-Ala), succinic acid

549 (Sigma-Aldrich, Taufkirchen, Germany) (referred to as succinate) or L-malic acid

550 (Merck Millipore, Darmstadt, Germany) (referred to as malate). For elevating the

- 551 extracellular availability of bicarbonate, 100 mM potassium HCO3 (Merck Millipore, 24

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

552 Darmstadt, Germany) was added to the CGXII-basis solution, which was sterile-

553 filtered before further use. In order to analyze the effect of different pH levels, either

554 HCl for lowering or KOH for increasing the pH was added to the CGXII-basis, which

555 was sterile-filtered afterwards. For cultivations in the presence of TCA cycle-

556 filling/refueling substrates, 50 mM citric acid monohydrate (citrate), 3 mM H-Glu-Ala-

557 OH dipeptide (Glu-Ala), 100 mM succinic acid (succinate) or 100 mM L-malic acid

558 (malate) was used. In experiments where gluconate or ribose were used as carbon

559 source, 100 mM D-gluconic acid sodium salt or 100 mM D-ribose were added,

560 respectively. Biomass formation was monitored during cultivation in shake flasks by

561 measuring OD600 or by measuring backscatter values during microtiter plate

562 cultivation (section 2.2). Where necessary, 25 µg/ml kanamycin was also

563 supplemented.

564 For the purpose of cloning and plasmid isolations, Escherichia coli DH5α strains were

565 grown in shake flasks in Lysogeny broth (LB) medium at 37°C directly inoculated

566 from a glycerol stock or from an LB-agar plate in shake flasks at 37°C. If necessary

567 for selection, 50 µg/ml kanamycin was also supplemented.

568 Microtiter plate cultivation

569 Online monitoring of growth and/or pH was performed in 48-well microtiter Flower

570 Plates® (m2p-labs GmbH, Baesweiler, Germany) sealed with sterile breathable

571 rayon films (VWR International, Pennsylvania, US) in a BioLector® microtiter

572 cultivation system (m2p-labs GmbH, Baesweiler, Germany) (61). The cultivation

573 conditions were adjusted as described earlier (62) and biomass formation was

574 recorded each 15 min as the backscattered light intensity (light wavelength 620 nm;

575 signal gain factor of 20) for 24 to 72 h at 30°C and 1,200 rpm. pH measurements

576 were performed with 48-well microtiter Flower Plates® being equipped with pH

577 optodes. Obtained data was evaluated using the software BioLection (m2p-labs, 25

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

578 Baesweiler, Germany) and Graph Pad Prism (Graphpad Prism 7 Software, Inc.,

579 California, USA).

580 Adaptive laboratory evolution

581 The adaptive laboratory evolution (ALE) of C. glutamicum ΔaceE and ΔaceE Δpyc

582 was performed in Deepwell-plates (VWR International, Pennsylvania, US) in a main

583 culture of 1 ml CGXII media containing 154 mM acetate and 222 mM glucose or

584 solely 154 mM acetate (as control without selection pressure) adjusted to an OD600 of

585 1. Cells were cultivated for 2-3 days before the next generation was inoculated

586 starting at an OD600 of 1 and cultivated again for 2-3 days. After each generation

587 step, glycerol stocks of cultures were prepared (20% glycerol) and stored at -80°C

588 allowing growth analysis and DNA sequencing of individual inoculations. In total, 16

589 serial transfers were analyzed.

590 Cloning techniques and recombinant DNA work

591 Standard molecular biology methods were performed according to J. Sambrook and

592 D. W. Russell (63). C. glutamicum ATCC 13032 chromosomal DNA was used as

593 template for PCR amplification of DNA fragments and prepared as described earlier

594 (64). DNA fragment and plasmid sequencing as well as synthesis of oligonucleotides

595 was performed by Eurofins Genomics (Ebersberg, Germany).

596 For the construction of plasmids (Table S2), DNA fragments were amplified using the

597 respective oligonucleotides (Table S1) and enzymatically assembled into a vector

598 backbone according to (65).

599 To achieve genomic deletion of pyc, ptsG, ppc and pgi, two-step homologous

600 recombination using the pK19mobsacB-system (66) was implemented. The suicide

601 plasmids (compare Table 2 and Table S2) were isolated from E. coli cells using the

602 QIAprep spin miniprep kit (Qiagen, Hilden, Germany). Electrocompetent C.

603 glutamicum ΔaceE and ΔaceE Δpyc cells were transformed with these plasmids by 26

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

604 electroporation (67). The first and second recombination events were performed and

605 verified as described in previous studies (68). The deletion of pyc, ptsG, ppc and pgi

606 was reviewed by amplification and sequencing using primers shown in Table S1.

607 Measurement of glucose concentrations

608 To measure the glucose concentration of the culture media at different time points,

609 cultivation was performed in 50 ml in shaking flasks. During cultivation, 0.5 ml

610 samples were taken every 3 h and centrifuged (16,000 x g). Supernatant was

611 collected and stored at -20°C until use.

612 The measurement of the actual glucose concentration was performed using D-

613 Glucose UV-Test Kit (r-biopharm, Darmstadt, Germany). Performance and

614 calculations according to manufacturer’s instructions were done, while absorption

615 was measured at 340 nm.

616 Further calculations concerning the glucose uptake rates could be done based on the

617 obtained glucose concentrations and OD600 values. According to (18), Formula 1 was

618 used:

619

620 Formula 1: Determination of glucose concentration in nmol gDW-1 h-1

푆 푛푚표푙 ∗ 푙−1 ∗ 푂퐷 −1 푛푚표푙 ( ) ∗ µ [( 600 ) ∗ ℎ−1] = [ ] −1 −1 푀 𝑔퐷푊 ∗ 푙 ∗ 푂퐷600 𝑔퐷푊 ∗ ℎ

621

-1 622 While S represents the slope of the glucose concentration versus OD600 [nmol * l *

-1 -1 - 623 OD600 ], M is the correlation between dry weight and OD600 [g dry weight * l * OD600

1 -1 624 ], and µ is the growth rate [h ]. According to A. Kabus et al. (69), an OD600 of 1

625 corresponds to 0.25 g dry weight l-1, so that this value was used as M throughout this

626 calculations.

27

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

627 Quantification of amino acid production

628 Using ultra-high performance liquid chromatography (uHPLC), amino acids were

629 quantified as ortho-phthaldialdehyde derivatives by automatic pre-column

630 derivatization. Separation of derivatives by reverse-phase chromatography was

631 performed on an Agilent 1290 Infinity LC ChemStation (Agilent, Santa Clara, USA)

632 equipped with a fluorescence detector. As eluent for the Zorbax Eclipse AAA 3.5

633 micron 4.6 x 7.5 mm column (Agilent, Santa Clara, USA), a gradient of Na-borate

634 buffer (10 mM Na2HPO4; 10 mM Na2B4O7, pH 8.2; adapted to operater’s guide) and

635 methanol was applied. Prior to analysis, samples were centrifuged for 10 min at

636 16,000 x g and 4°C and diluted 1:100.

637 Monitoring of cellular proliferation by cell staining

638 For staining of proliferating cells, the PKH67 green fluorescent cell linker kit for

639 general cell membrane labeling (Sigma Aldrich, Munich, Germany) was used and the

640 protocol was adapted to O. Kotte et al. (36). From an exponentially growing pre-

641 culture in CGXII minimal medium containing 222 mM glucose and 154 mM acetate,

642 1.5 x 109 cells were harvested by centrifugation for four minutes at 4000 g and 4°C.

643 Then, the cells were washed again in 5 ml ice-cold CGXII basic solution, without

644 MgSO4, CaCl2, biotin, trace elements and protocatechuic acid. For staining, the cell

645 pellet was resuspended in 500 µl dilution buffer C (Sigma Aldrich, Munich, Germany)

646 at room temperature and a freshly prepared mixture of 10 µl PKH67 dye (Sigma

647 Aldrich, Munich, Germany) and 500 µl dilution buffer C was added. Subsequently,

648 cells were incubated for three minutes at room temperature and afterwards, 4 ml ice-

649 cold filtered CGXII basic solution containing 1% (w/v) bovine serum albumin, 1 mM

650 MgSO4 and 0.1 mM CaCl2 were added. Then, the cells were centrifuged for four

651 minutes at 4000 g and 4°C and the cell pellet was washed twice. Finally, the cells

28

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

652 were resuspended in CGXII minimal medium containing 222 mM glucose and 154

653 mM acetate and cultivated in a microtiter cultivation system.

654 Flow cytometry

655 Flow cytometric (FC) analyzes were conducted on a FACSAria II flow cytometer

656 (Becton Dickinson, San Jose, USA) equipped with a blue solid state laser (488 nm

657 excitation). Forward-scatter characteristics (FSC) and side-scatter characteristics

658 (SSC) were recorded as small-angle and orthogonal scatters of the 488-nm laser,

659 respectively. PKH67 fluorescence was detected using a 502-nm long-pass and a

660 530/30-nm band-pass filter set. FACS-Diva software 6.0 was used to monitor the

661 measurements. During analyzes, thresholding on FSC was applied to remove

662 background noise. For FC analyzes, PKH67-stained culture samples were diluted to

663 an OD600 of 0.05 in FACSFlow™ sheath fluid buffer (BD, Heidelberg, Germany). The

664 analysis software FlowJo V.10.0.8 was used to visualize and evaluate the data (Tree

665 Star, Ashland, USA).

666 DNA microarrays

667 For analysis of the transcriptome, C. glutamicum ΔaceE and ΔaceE Δpyc were

668 cultivated in triplicates as described in 2.1 in 50 ml CGXII containing 154 mM acetate

669 and 222 mM glucose in shake flasks. After reaching exponential phase at an OD600

670 of around 12-15, the cell suspension was harvested by centrifugation (4256 g, 10

671 min, 4°C). The resulting pellet was directly frozen in liquid nitrogen and stored at -

672 80°C. RNA preparation, cDNA synthesis as well as microchip hybridization, scanning

673 and evaluation were performed as described by previous studies (70).

674 GC-ToF-MS analysis

675 For analysis of the metabolome, samples of C. glutamicum ΔaceE in exponential

676 phase as well as ΔaceE Δpyc in stationary and exponential phase were taken and

677 cell disruption was accomplished using hot methanol. Further sample preparation, 29

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

678 derivatization, MS operation and peak identification was accomplished according to

679 N. Paczia et al. (71) in an Agilent 6890N gas chromatograph (Agilent, Santa Clara,

680 USA) coupled to a Waters Micromass GCT Premier high resolution time-of-flight

681 mass spectrometer (Waters, Milford, USA). Identification of known metabolites was

682 accomplished using in-house database JuPoD, the commercial database NIST17

683 (National Institute of Standards and Technology, US) and the database GMD (MPI of

684 Molecular Plant Physiology, Golm) (72).

685 Whole-genome sequencing

686 In order to sequence the whole genome of C. glutamicum ΔaceE Δpyc mutants from

687 the ALE experiment using next-generation sequencing (NGS), genomic DNA was

688 prepared using the NucleoSpin® Microbial DNA Kit (Macherey-Nagel, Düren,

689 Germany) according to manufacturer’s instructions. Consequently, the concentrations

690 of the purified genomic DNA were measured using Qubit Fluorometer 2.0 (Invitrogen,

691 California, USA) according to manufacturer’s instructions. Overall, 4 µg purified

692 genomic DNA was employed for the preparation of genome sequencing using

693 TruSeq® DNA Library Prep Kit and MiSeq Reagent Kit v1 (Illumia, San Diego, USA)

694 according to manufacturer’s instructions. The sequencing run was performed in a

695 MiSeq system (Illumina, San Diego, USA). Data analysis and base calling was

696 performed with the Illumina instrument software. Consequent fastq output files were

697 examined using CLC Genomics Workbench 9 (Qiagen Aarhus, Denmark).

698

699 Acknowledgements

700 We thank Jochem Gätgens for performing GC-ToF analyses, as well as Bastian

701 Blombach and Stephan Noack for fruitful discussions. The authors acknowledge

702 financial support by the Helmholtz Association (grant W2/W3-096).

703 30

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

704 References

705 1. Bailey JE, Ollis DF. 1986. Biochemical Engineering Fundamentals. NY:

706 McGraw-Hill, New York.

707 2. Blombach B, Takors R. 2015. CO2 – Intrinsic Product, Essential Substrate,

708 and Regulatory Trigger of Microbial and Mammalian Production Processes.

709 Front Bioeng Biotechnol 3.

710 3. de Kok A, Hengeveld AF, Martin A, Westphal AH. 1998. The pyruvate

711 dehydrogenase multi-enzyme complex from Gram-negative bacteria. Biochim

712 Biophys 1385:353-366.

713 4. Eikmanns BJ, Blombach B. 2014. The pyruvate dehydrogenase complex of

714 Corynebacterium glutamicum: an attractive target for metabolic engineering. J

715 Biotechnol 192 Pt B:339-45.

716 5. Schreiner ME, Fiur D, Holátko J, Pátek M, Eikmanns BJ. 2005. E1 Enzyme of

717 the Pyruvate Dehydrogenase Complex in Corynebacterium glutamicum :

718 Molecular Analysis of the Gene and Phylogenetic Aspects. J Bacteriol

719 187:6005-6018.

720 6. Becker J, Rohles CM, Wittmann C. 2018. Metabolically engineered

721 Corynebacterium glutamicum for bio-based production of chemicals, fuels,

722 materials, and healthcare products. Metab Eng 50:122-141.

723 7. Kogure T, Inui M. 2018. Recent advances in metabolic engineering of

724 Corynebacterium glutamicum for bioproduction of value-added aromatic

725 chemicals and natural products. Appl Microbiol Biotechnol 102:8685-8705.

726 8. Wang X, Zhang H, Quinn PJ. 2018. Production of L-valine from metabolically

727 engineered Corynebacterium glutamicum. Appl Microbiol Biotechnol

728 102:4319-4330.

31

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

729 9. Wendisch VF, Mindt M, Perez-Garcia F. 2018. Biotechnological production of

730 mono- and diamines using bacteria: recent progress, applications, and

731 perspectives. Appl Microbiol Biotechnol 102:3583-3594.

732 10. Eggeling L, Bott M. 2005. Handbook of Corynebacterium glutamicum.

733 Academic Press, Inc., Boca Raton, FL.

734 11. Blombach B, Schreiner ME, Bartek T, Oldiges M, Eikmanns BJ. 2008.

735 Corynebacterium glutamicum tailored for high-yield L-valine production. Appl

736 Microbiol Biotechnol 79:471-9.

737 12. Blombach B, Schreiner ME, Holatko J, Bartek T, Oldiges M, Eikmanns BJ.

738 2007. L-valine production with pyruvate dehydrogenase complex-deficient

739 Corynebacterium glutamicum. Appl Environ Microbiol 73:2079-84.

740 13. Buchholz J, Schwentner A, Brunnenkan B, Gabris C, Grimm S, Gerstmeir R,

741 Takors R, Eikmanns BJ, Blombach B. 2013. Platform engineering of

742 Corynebacterium glutamicum with reduced pyruvate dehydrogenase complex

743 activity for improved production of L-lysine, L-valine, and 2-ketoisovalerate.

744 Appl Environ Microbiol 79:5566-75.

745 14. Cocaign M, Monnet C, Lindley ND. 1993. Batch kinetics of Corynebacterium

746 glutamicum during growth on various carbon substrates: use of substrate

747 mixtures to localise metabolic bottlenecks. Appl Microbiol Biotechnol 40:526-

748 530.

749 15. Wendisch VF, de Graaf AA, Sahm H, Eikmanns BJ. 2000. Quantitative

750 Determination of Metabolic Fluxes during Coutilization of Two Carbon

751 Sources: Comparative Analyses with Corynebacterium glutamicum during

752 Growth on Acetate and/or Glucose. J Bacteriol 182:3088-3096.

32

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

753 16. Dominguez H, Cocaign-Bousquet M, Lindley ND. 1997. Simultaneous

754 consumption of glucose and fructose from sugar mixtures during batch growth

755 of Corynebacterium glutamicum. Appl Microbiol Biotechnol 47:600-603.

756 17. Stansen C, Uy D, Delaunay S, Eggeling L, Goergen J-L, Wendisch VF. 2005.

757 Characterization of a Corynebacterium glutamicum lactate utilization operon

758 induced during temperature-triggered glutamate production. Appl Environ

759 Microbiol 71:5920-5928.

760 18. Frunzke J, Engels V, Hasenbein S, Gätgens C, Bott M. 2008. Co-ordinated

761 regulation of gluconate catabolism and glucose uptake in Corynebacterium

762 glutamicum by two functionally equivalent transcriptional regulators, GntR1

763 and GntR2. Mol Microbiol 67:305-322.

764 19. Cozzone AJ. 1998. Regulation of Acetate Metabolism by Protein

765 Phosphorylation in Enteric Bacteria. Annu Rev Microbiol 52:127-164.

766 20. Cronan J, P. Clark D. 2005. Two-Carbon Compounds and Fatty Acids as

767 Carbon Sources. EcoSal Plus 1.

768 21. Sauer U, Eikmanns BJ. 2005. The PEP-pyruvate-oxaloacetate node as the

769 switch point for carbon flux distribution in bacteria. FEMS Microbiol Rev

770 29:765-94.

771 22. Eikmanns BJ, Follettie MT, Griot MU, Sinskey AJ. 1989. The

772 phosphoenolpyruvate carboxylase gene of Corynebacterium glutamicum:

773 Molecular cloning, nucleotide sequence, and expression. Mol Gen Genet

774 218:330-339.

775 23. O'Regan M, Thierbach G, Bachmann B, Villeval D, Lepage P, Viret JF,

776 Lemoine Y. 1989. Cloning and nucleotide sequence of the

777 phosphoenolpyruvate carboxylase-coding gene of Corynebacterium

778 glutamicum ATCC13032. Gene 77:237-251. 33

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

779 24. Peters-Wendisch PG, Kreutzer C, Kalinowski J, Patek M, Sahm H, Eikmanns

780 BJ. 1998. Pyruvate carboxylase from Corynebacterium glutamicum:

781 characterization, expression and inactivation of the pyc gene. Microbiology

782 144 (Pt 4):915-27.

783 25. Kai Y, Matsumura H, Inoue T, Terada K, Nagara Y, Yoshinaga T, Kihara A,

784 Tsumura K, Izui K. 1999. Three-dimensional structure of phosphoenolpyruvate

785 carboxylase: a proposed mechanism for allosteric inhibition. Proceedings of

786 the National Academy of Sciences of the United States of America 96:823-

787 828.

788 26. Repaske R, Repaske AC, Mayer RD. 1974. Carbon dioxide control of lag

789 period and growth of Streptococcus sanguis. J Bacteriol 117:652-659.

790 27. Talley RS, Baugh CL. 1975. Effects of bicarbonate on growth of Neisseria

791 gonorrhoeae: replacement of gaseous CO2 atmosphere. Appl Microbiol

792 29:469-471.

793 28. Bäumchen C, Knoll A, Husemann B, Seletzky J, Maier B, Dietrich C,

794 Amoabediny G, Buchs J. 2007. Effect of elevated dissolved carbon dioxide

795 concentrations on growth of Corynebacterium glutamicum on D-glucose and

796 L-lactate. J Biotechnol 128:868-74.

797 29. Blombach B, Buchholz J, Busche T, Kalinowski J, Takors R. 2013. Impact of

- 798 different CO2/HCO3 levels on metabolism and regulation in Corynebacterium

799 glutamicum. J Biotechnol 168:331-40.

800 30. Bartek T, Blombach B, Lang S, Eikmanns BJ, Wiechert W, Oldiges M, Noh K,

801 Noack S. 2011. Comparative 13C metabolic flux analysis of pyruvate

802 dehydrogenase complex-deficient, L-valine-producing Corynebacterium

803 glutamicum. Appl Environ Microbiol 77:6644-52.

34

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

804 31. Dominguez H, Lindley ND. 1996. Complete Sucrose Metabolism Requires

805 Fructose Phosphotransferase Activity in Corynebacterium glutamicum To

806 Ensure Phosphorylation of Liberated Fructose. Appl Environ Microbiol

807 62:3878-80.

808 32. Moon MW, Kim HJ, Oh TK, Shin CS, Lee JS, Kim SJ, Lee JK. 2005. Analyses

809 of enzyme II gene mutants for sugar transport and heterologous expression of

810 fructokinase gene in Corynebacterium glutamicum ATCC 13032. FEMS

811 Microbiol Lett 244:259-66.

812 33. Peters-Wendisch PG, Eikmanns BJ, Thierbach G, Bachmann B, Sahm H.

813 1993. Phosphoenolpyruvate carboxylase in Corynebacterium glutamicum is

814 dispensable for growth and lysine production. FEMS Microbiol Lett 112:269-

815 274.

816 34. Mahr R, Gätgens C, Gätgens J, Polen T, Kalinowski J, Frunzke J. 2015.

817 Biosensor-driven adaptive laboratory evolution of l-valine production in

818 Corynebacterium glutamicum. Metab Eng 32:184-94.

819 35. Reinscheid DJ, Eikmanns BJ, Sahm H. 1994. Characterization of the isocitrate

820 lyase gene from Corynebacterium glutamicum and biochemical analysis of the

821 enzyme. J Bacteriol 176:3474-3483.

822 36. Kotte O, Volkmer B, Radzikowski JL, Heinemann M. 2014. Phenotypic

823 bistability in Escherichia coli's central carbon metabolism. Mol Syst Biol

824 10:736.

825 37. Takhaveev V, Heinemann M. 2018. Metabolic heterogeneity in clonal microbial

826 populations. Curr Opin Microbiol 45:30-38.

827 38. Heinemann M, Zenobi R. 2011. Single cell metabolomics. Curr Opin

828 Biotechnol 22:26-31.

35

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

829 39. Davis KM, Isberg RR. 2016. Defining heterogeneity within bacterial

830 populations via single cell approaches. Bioessays 38:782-90.

831 40. Ikeda M, Mizuno Y, Awane SI, Hayashi M, Mitsuhashi S, Takeno S. 2011.

832 Identification and application of a different glucose uptake system that

833 functions as an alternative to the phosphotransferase system in

834 Corynebacterium glutamicum. Appl Microbiol Biotechnol 90:1443-1451.

835 41. Lindner SN, Seibold GM, Henrich A, Krämer R, Wendisch VF. 2011.

836 Phosphotransferase system-independent glucose utilization in

837 Corynebacterium glutamicum by inositol permeases and glucokinases. Appl

838 Environ Microbiol 77:3571-3581.

839 42. Peters-Wendisch PG, Schiel B, Wendisch VF, Katsoulidis E, Mockel B, Sahm

840 H, Eikmanns BJ. 2001. Pyruvate carboxylase is a major bottleneck for

841 glutamate and lysine production by Corynebacterium glutamicum. J Mol

842 Microbiol Biotechnol 3:295-300.

843 43. Kortmann M, Mack C, Baumgart M, Bott M. 2019. Pyruvate Carboxylase

844 Variants Enabling Improved Lysine Production from Glucose Identified by

845 Biosensor-Based High-Throughput Fluorescence-Activated Cell Sorting

846 Screening. ACS Synth Biol 8:274-281.

847 44. Chatterjee R, Millard CS, Champion K, Clark DP, Donnelly MI. 2001. Mutation

848 of the ptsG gene results in increased production of succinate in fermentation

849 of glucose by Escherichia coli. Appl Environ Microbiol 67:148-154.

850 45. Wang Q, Wu C, Chen T, Chen X, Zhao X. 2006. Expression of galactose

851 permease and pyruvate carboxylase in Escherichia coli ptsG mutant increases

852 the growth rate and succinate yield under anaerobic conditions. Biotechnol

853 Lett 28:89-93.

36

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

854 46. Vanderpool CK, Gottesman S. 2007. The novel transcription factor SgrR

855 coordinates the response to glucose-phosphate stress. J Bacteriol 189:2238-

856 48.

857 47. Morita T, El-Kazzaz W, Tanaka Y, Inada T, Aiba H. 2003. Accumulation of

858 glucose 6-phosphate or fructose 6-phosphate is responsible for destabilization

859 of glucose transporter mRNA in Escherichia coli. J Biol Chem 278:15608-14.

860 48. Kadner RJ, Murphy GP, Stephens CM. 1992. Two mechanisms for growth

861 inhibition by elevated transport of sugar phosphates in Escherichia coli. J Gen

862 Microbiol 138:2007-14.

863 49. Kessler JR, Cobe BL, Richards GR. 2017. Stringent Response Regulators

864 Contribute to Recovery from Glucose Phosphate Stress in Escherichia coli.

865 Appl Environ Microbiol 83.

866 50. Vanderpool CK, Gottesman S. 2004. Involvement of a novel transcriptional

867 activator and small RNA in post-transcriptional regulation of the glucose

868 phosphoenolpyruvate phosphotransferase system. Mol Microbiol 54:1076-89.

869 51. Engels V, Wendisch VF. 2007. The DeoR-type regulator SugR represses

870 expression of ptsG in Corynebacterium glutamicum. J Bacteriol 189:2955-66.

871 52. Gaigalat L, Schlüter JP, Hartmann M, Mormann S, Tauch A, Pühler A,

872 Kalinowski J. 2007. The DeoR-type transcriptional regulator SugR acts as a

873 repressor for genes encoding the phosphoenolpyruvate:sugar

874 phosphotransferase system (PTS) in Corynebacterium glutamicum. BMC Mol

875 Biol 8:104.

876 53. Tanaka Y, Teramoto H, Inui M, Yukawa H. 2008. Regulation of expression of

877 general components of the phosphoenolpyruvate: carbohydrate

878 phosphotransferase system (PTS) by the global regulator SugR in

879 Corynebacterium glutamicum. Appl Microbiol Biotechnol 78:309-18. 37

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

880 54. Blombach B, Arndt A, Auchter M, Eikmanns BJ. 2009. L-valine production

881 during growth of pyruvate dehydrogenase complex-deficient Corynebacterium

882 glutamicum in the presence of ethanol or by inactivation of the transcriptional

883 regulator SugR. Appl Environ Microbiol 75:1197-200.

884 55. Repaske R, Clayton MA. 1978. Control of Escherichia coli growth by CO2. J

885 Bacteriol 135:1162-4.

886 56. Valley G, Rettger LF. 1927. The Influence of Carbon Dioxide on Bacteria. J

887 Bacteriol 14:101-37.

888 57. Bracher JM, Martinez-Rodriguez OA, Dekker WJC, Verhoeven MD, van Maris

889 AJA, Pronk JT. 2019. Reassessment of requirements for anaerobic xylose

890 fermentation by engineered, non-evolved strains.

891 FEMS Yeast Res 19.

892 58. Gonzalez JE, Long CP, Antoniewicz MR. 2017. Comprehensive analysis of

893 glucose and xylose metabolism in Escherichia coli under aerobic and

894 anaerobic conditions by (13)C metabolic flux analysis. Metab Eng 39:9-18.

895 59. Kalinowski J, Bathe B, Bartels D, Bischoff N, Bott M, Burkovski A, Dusch N,

896 Eggeling L, Eikmanns BJ, Gaigalat L, Goesmann A, Hartmann M, Huthmacher

897 K, Krämer R, Linke B, McHardy AC, Meyer F, Möckel B, Pfefferle W, Pühler A,

898 Rey DA, Rückert C, Rupp O, Sahm H, Wendisch VF, Wiegräbe I, Tauch A.

899 2003. The complete Corynebacterium glutamicum ATCC 13032 genome

900 sequence and its impact on the production of L-aspartate-derived amino acids

901 and vitamins. J Biotechnol 104:5-25.

902 60. Keilhauer C, Eggeling L, Sahm H. 1993. Isoleucine synthesis in

903 Corynebacterium glutamicum: Molecular analysis of the ilvB-ilvN-ilvC operon.

904 J Bacteriol 175:5595-5603.

38

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

905 61. Kensy F, Zang E, Faulhammer C, Tan RK, Büchs J. 2009. Validation of a

906 high-throughput fermentation system based on online monitoring of biomass

907 and fluorescence in continuously shaken microtiter plates. Microbial Cell

908 Factories 8:1-17.

909 62. Mustafi N, Gr A, Kohlheyer D, Bott M, Frunzke J. 2012. The development and

910 application of a single-cell biosensor for the detection of L-methionine and

911 branched-chain amino acids. Metabolic Engineering 14:449-457.

912 63. Sambrook J, Russell DW. 2001. Molecular Cloning: A Laboratory Manual, 3

913 ed. Cold Spring Harbor Laboratory Press.

914 64. Eikmanns BJ, Thum-Schmitz N, Eggeling L, Lüdtke KU, Sahm H. 1994.

915 Nucleotide sequence, expression and transcriptional analysis of the

916 Corynebacterium glutamicum gltA gene encoding .

917 Microbiology 140:1817-1828.

918 65. Gibson DG, Young L, Chuang RY, Venter JC, Hutchison CA, Smith HO. 2009.

919 Enzymatic assembly of DNA molecules up to several hundred kilobases.

920 Nature Methods 6:343-345.

921 66. Schäfer A, Tauch A, Jäger W, Kalinowski J, Thierbach G, Pühler A. 1994.

922 Small mobilizable multi-purpose cloning vectors derived from the Escherichia

923 coli plasmids pK18 and pK19: selection of defined deletions in the

924 chromosome of Corynebacterium glutamicum. Gene 145:69-73.

925 67. Van der Rest ME, Lange C, Molenaar D. 1999. A heat shock following

926 electroporation induces highly efficient transformation of Corynebacterium

927 glutamicum with xenogeneic plasmid DNA. Appl Microbiol Biotechnol 52:541-

928 545.

39

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

929 68. Niebisch A, Bott M. 2001. Molecular analysis of the cytochrome bc1-

930 aa3branch of the Corynebacterium glutamicum respiratory chain containing an

931 unusual diheme cytochrome c1. Arch Microbiol 175:282-294.

932 69. Kabus A, Georgi T, Wendisch VF, Bott M. 2007. Expression of the Escherichia

933 coli pntAB genes encoding a membrane-bound transhydrogenase in

934 Corynebacterium glutamicum improves L-lysine formation. Appl Microbiol

935 Biotechnol 75:47-53.

936 70. Baumgart M, Luder K, Grover S, Gätgens C, Besra GS, Frunzke J. 2013.

937 IpsA, a novel LacI-type regulator, is required for inositol-derived lipid formation

938 in Corynebacteria and Mycobacteria. BMC Biology 11.

939 71. Paczia N, Nilgen A, Lehmann T, Gätgens J, Wiechert W, Noack S. 2012.

940 Extensive exometabolome analysis reveals extended overflow metabolism in

941 various microorganisms. Microbial Cell Factories 11:1-14.

942 72. Hummel J, Strehmel N, Selbig J, Walther D, Kopka J. 2010. Decision tree

943 supported substructure prediction of metabolites from GC-MS profiles.

944 Metabolomics 6:322-333.

945 73. Kinoshita S, Udaka S, Shimono M. 2004. Studies on the amino acid

946 fermentation. Part 1. Production of L-glutamic acid by various microorganisms.

947 J Gen Appl Microbiol 50:331-343.

948 74. Schwentner A, Feith A, Münch E, Busche T, Rückert C, Kalinowski J, Takors

949 R, Blombach B. 2018. Metabolic engineering to guide evolution - Creating a

950 novel mode for L-valine production with Corynebacterium glutamicum. Metab

951 Eng 47:31-41.

952 75. Cremer J, Eggeling L, Sahm H. 1990. Cloning the dapA dapB cluster of the

953 lysine-secreting bacterium Corynebacterium glutamicum. Mol Gen Genet

954 220:478-480. 40

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

955

41

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

956 Tables

957 Table 1: Bacterial strains used in this study.

Strain Characteristics Reference C. glutamicum Wild type, biotin-auxotroph S. Kinoshita et al. (73) ATCC 13032 ∆aceE ATCC 13032 derivate with deletion of aceE M. E. Schreiner et al. (5) ∆aceE ∆pyc ATCC 13032 derivate with deletion of aceE This work and pyc ∆aceE ∆ptsG ATCC 13032 derivate with deletion of aceE This work and ptsG ∆aceE ∆pyc ∆ptsG ATCC 13032 derivate with deletion of aceE This work and pyc and ptsG ∆aceE ∆pgi ATCC 13032 derivate with deletion of aceE This work and pgi ∆aceE ∆pyc ∆pgi ATCC 13032 derivate with deletion of aceE This work and pyc and pgi ∆aceE ∆ppc ATCC 13032 derivate with deletion of aceE This work and ppc ∆aceE ∆pyc ∆ppc ATCC 13032 derivate with deletion of aceE, This work pyc and ppc ∆aceE ureD- ATCC 13032 derivate with deletion of aceE R. Mahr et al. (34) E188* and ureD-E188* (Glu188 to stop codon) ∆aceE ureD- ATCC 13032 derivate with deletion of aceE, This work E188* ∆pyc pyc and ureD-E188* (Glu188 to stop codon) ∆pyc ATCC 13032 derivate with deletion of pyc P. G. Peters- Wendisch et al. (24) ∆pyc ∆ppc ATCC 13032 derivate with deletion of aceE A. Schwentner et al. and ppc (74) E. coli DH5α fhuA2 lac(del)U169 phoA glnV44 Φ80' Invitrogen lacZ(del)M15 gyrA96 recA1 relA1 endA1 thi- 1 hsdR17; for general cloning purposes 958

959

42

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

960 Table 2: Plasmids used in this study.

Plasmid Characteristics Reference pK19mobsacB Contains negative (sacB) and positive (kanR) selection A. Schäfer marker for genomic integration and deletion et al. (66) pK19mobsacB- KanR; derivative of pK19mobsacB for partial pyc gene P. G. ∆pyc deletion, contains only the first 499 bp and last 493 bp of Peters- pyc, for the deletion of 2429 bp in the middle Wendisch et al. (24) pK19mobsacB- KanR; derivative of pK19mobsacB for partial ptsG gene This work ∆ptsG deletion, contains only the last 491 bp of ptsG, for the deletion of the first 1561 bp pK19mobsacB- KanR; derivative of pK19mobsacB for pgi gene deletion, This work ∆pgi contains only the last 537 bp of pgi, for the deletion of the

first 1086 bp pK19mobsacB- KanR; derivative of pK19mobsacB for partial ppc gene J. Buchholz ∆ppc deletion, contains only the first 391 bp and last 447 bp of et al. (13) ppc, for the deletion of 1922 bp in the middle pK19mobsacB- KanR; derivative of pK19mobsacB for ureD gene deletion, R. Mahr et ureD-E188* exchange of Glu188 to stop codon in ureD al. (34)

R pJC1 Kan , oriVEc, oriVCg; E. coli - C. glutamicum shuttle vector J. Cremer et al. (75)

pJC1-venus- KanR; derivative of pJC1, contains terminator sequence of M. term-BS Bacillus subtilis behind venus Baumgart et al. (70) pJC1-pyc KanR; derivative of pJC1-venus-term-BS, contains pyc This work under control of its native promoter Ppyc

R pJC1-lacI/Ptac- Kan ; derivative of pJC1-venus-term-BS, contains aceA This work aceA-linker- followed by a linking sequence (5’- aceB ACTAGAAATAATTTTGTTTAACTTTAAGAAGGAGATATACAT-

3’) and aceB under control of the inducible promoter Ptac R pJC1-lacI/Ptac- Kan ; derivative of pJC1-venus-term-BS, contains aceB This work aceB-linker- followed by a linking sequence (5’- aceA ACTAGAAATAATTTTGTTTAACTTTAAGAAGGAGATATACAT-

3’) and aceA under control of the inducible promoter Ptac 961

962

43

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

963 Table 3: Glucose consumption rate of C. glutamicum PDHC-deficient strains in the

964 exponential growth phase. For an overview of glucose consumption of the entire time

965 course of the experiment including the stationary phase, the reader is referred to Fig.

966 S2.

Strain Glucose consumption rate (nmol min-1 g-1)a ΔaceE 16.81 ± 3.01 ΔaceE Δpyc 10.87 ± 4.95 ΔaceE ΔptsG 7.24 ± 3.97 ΔaceE Δpyc ΔptsG 7.32 ± 1.88 967

968 a Means ± standard deviation of three independent biological replicates.

969

970

44

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

971 Table 4: Overview of lag phases and growth rates of C. glutamicum ΔaceE and

972 ΔaceE Δpyc cultivated on different TCA cycle carbon sources in CGXII medium

973 containing 154 mM acetate and 222 mM glucose.

ΔaceE ΔaceE Δpyc Added carbon source* Lag Growth rate (h- Lag Growth Δt (%)** Δt (%)** phase 1) phase rate (h-1) 0.27 ± / - 0 0.41 ± 0.01 + 100 0.02 0.33 ± Glu-Ala (3 mM) - 0 0.35 ± 0.01 ÷ 10 0.01 0.35 ± Citrate (50 mM) - 0 0.39 ± 0.01 ÷ 40 0.01 0.34 ± Succinate (100 mM) - -8 0.36 ± 0.01 ÷ 15 0.02 0.35 ± Malate (100 mM) - -8 0.35 ± 0.01 ÷ 38 0.02 974

975 *Additionally the medium contained 154 mM acetate and 222 mM glucose,

976 ** The difference between the first doubling time of strain ΔaceE and ΔaceE Δpyc

977 was defined as 100%. + = elongated lag phase referring to growth on 154 mM

978 acetate with 222 mM glucose, ÷ = shorter lag phase, - = no lag phase.

979

980

45

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

981 Table 5: Overview of lag phases and growth rates of C. glutamicum Δpyc, Δpyc Δppc

982 and ΔaceE Δpyc Δppc cultivated on different TCA cycle carbon sources in CGXII

983 medium containing 154 mM acetate and 222 mM glucose.

Δpyc Δpyc Δppc ΔaceE Δpyc Δppc Added carbon Growth rate Growth rate Growth rate source* Δt ** Δt ** Δt ** lag (h-1)*** lag (h-1)*** lag (h-1)*** / 6 h 0.17 ± 0.02 No growth No growth

Glu-Ala n.d. n.d. 8 h 0.21 ± 0.01 30 h 0.14 ± 0.01 (3 mM) Citrate 3 h 0.19 ± 0.01 5 h 0.26 ± 0.01 8 h 0.20 ± 0.03 (50 mM) Succinate 2 h 0.2 ± 0.02 3 h 0.16 ± 0.03 5 h 0.12 ± 0.03 (100 mM) Malate 3 h 0.21 ± 0.01 17 h 0.18 ± 0.01 18 h 0.24 ± 0.01 (100 mM) 984

985 *Additionally the medium contained 154 mM acetate and 222 mM glucose,

986 ** = Δtlag represents the difference in lag phase duration (h) between ΔaceE growing

987 on glucose-acetate mixture and the strain of interest growing on the respective

988 carbon source(s).

989 *** = for comparison, growth rate in glucose-acetate mixture was 0.41 h-1 for ΔaceE

-1 990 and 0.28 h for ΔaceE Δpyc in this experiment. n.d. = not determined

991

992

46

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

993 Table 6: Comparative transcriptome analysis of C. glutamicum ΔaceE Δpyc and

994 strain ΔaceE and during growth on glucose and acetate. Shown are selected genes

995 encoding for central metabolic enzymes or regulators (for mRNA ratios of ΔaceE

996 Δpyc versus ΔaceE ≤ 0.70 values are shaded in red; mRNA ratio ≥ 1.70 are shaded

997 in green). p-values >0.05 are highlighted red (n=3 biological replicates). For the

998 complete dataset, the reader is referred to Table S3.

Gene Gene Annotation mRNA p-Value name ratio PTS-components cg1537 ptsG glucose-specific enzyme II BC component of PTS 0.59 0.02 cg2117 ptsI PEP:sugar phosphotransferase system enzyme I 0.67 0.01 cg2121 ptsH phosphocarrier protein HPR of PTS 1.16 0.11

Myo-Inositol transporter

cg0223 iolT1 myo-Inositol transporter 1.97 0.04 cg3387 iolT2 myo-Inositol transporter 1.41 0.19

Transcriptional regulators of ptsG

cg2783 gntR1 transcriptional regulator, GntR family 0.76 0.05 cg1935 gntR2 transcriptional regulator, GntR family 0.97 0.46 cg2115 sugR DeoR-type transcriptional regulator of ptsG, ptsS and 0.79 0.03 ptsF

Acetate metabolism

cg3047 ackA acetate/propionate kinase 1.99 0.01 cg3048 pta phosphate acetyltransferase 1.34 0.04 cg2560 aceA isocitrate lyase 1.70 0.12

cg2559 aceB malate synthase 1.88 0.02

Anaplerosis

cg1787 ppc phosphoenolpyruvate carboxylase 0.91 0.32 cg3169 pck phosphoenolpyruvate carboxykinase (GTP) 0.72 0.03 cg1458 odx oxaloacetate decarboxylase 1.07 0.32 cg3335 malE malic enzyme 0.48 0.26

Further enzymes of central carbon metabolism

cg0973 pgi glucose-6-phosphate isomerase 0.44 0.03

47

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

cg2291 pyk pyruvate kinase 0.59 0.01 cg2891 pqo pyruvate chinon oxidoreductase 0.47 0.02 999

48

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

1000 Table 7: Key mutations identified in adaptive laboratory evolution experiment of

1001 C. glutamicum ΔaceE Δpyc.

Single clone Batch No. Gene Gene Annotation Mutation ΔaceE Δpyc ID name 1 7 cg2117 ptsI EI enzyme, general component Exchange of PTS (EC: 2.7.3.9) R132H 2 8 cg0601 rpsC 30 S ribosomal protein S3 Exchange R225H 3 10 cg2117 ptsI EI enzyme, general component Exchange of PTS (EC: 2.7.3.9) G15S

1002

1003

49

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

1004 Figure Legends

1005

1006 1007 Figure 1: Schematic overview of the central carbon metabolism of C.

1008 glutamicum. The main aspects of glycolysis, gluconeogenesis, pentose phosphate

1009 pathway, tricarboxylic acid cycle (TCA) and anaplerosis are shown, while grey boxes

1010 represent relevant enzymes. Carboxylation as well as decarboxylation steps are

1011 given. ABCRib = ATP-binding cassette transporter for ribose, AK = acetate kinase, CA

1012 = carbonic anhydrase, , GntP = gluconate permease, GPDH = glucose-6P

1013 dehydrogenase, ICD = , KGDH = α-ketoglutarate

1014 dehydrogenase complex, PCx = pyruvate carboxylase, PDHC = pyruvate

1015 dehydrogenase complex, PEPCk = PEP carboxykinase, PEPCx = PEP carboxylase,

1016 PGDH = 6-P gluconate dehydrogenase, PGI = phosphoglucose isomerase, PK =

50

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

1017 pyruvate kinase, PQO = pyruvate:quinone oxidoreductase, PTA =

1018 phosphotransacetylase, PTSGlc = permease of phosphotransferase system for

1019 glucose, PTSFru= permease of phosphotransferase system for fructose.

1020

51

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

1021 1022 Figure 2: Impact of glucose on aceE and pyc mutant strains. Growth curves

1023 shown are based on the backscatter measurements in a microtiter cultivation system,

1024 inoculated at an OD600 of 1. Data represent the average of three biological replicates,

1025 error bars the standard deviation. (A) The C. glutamicum strains ∆aceE and ∆aceE

1026 ∆pyc, as well as the wild type were inoculated in CGXII media containing 154 mM

1027 acetate (left) or 154 mM acetate plus 222 mM glucose (right). (B) Strains ΔaceE and

1028 ΔaceE Δpyc as well as WT cells were cultivated in CGXII medium containing 154 mM

52

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

1029 acetate and 222 mM glucose and supplemented either with (red lines) or without 100

- 1030 mM KHCO3 (blue and black lines). The box zooms in the time interval between 4 h

1031 and 15 h. (C) Strains ΔaceE Δpyc (left) and Δpyc (right) were transformed with pJC1-

1032 pyc plasmid for complementation of the pyc deletion or with pJC1-empty vector as

1033 control. Cultures were inoculated in CGXII containing 154 mM acetate, 222 mM

1034 glucose and 25 µg/ml kanamycin.

1035

53

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

1036 1037 Figure 3: Influence of glucose consumption on the growth of C. glutamicum

1038 ΔaceE and ΔaceE Δpyc. Growth curves shown are based on the backscatter

1039 measurements in a microtiter cultivation system, inoculated at an OD600 of 1. (A) The

1040 strains ∆aceE (left) and ∆aceE ∆pyc (right) were inoculated to an OD600 of 1 in CGXII

1041 medium containing 154 mM acetate and either 0 mM (red), 100 mM (grey), 222 mM

1042 (blue) or 250 mM glucose (orange). (B) Deletion of the ptsG gene in the ΔaceE and

1043 ΔaceE Δpyc strains restored growth on glucose acetate mixtures (shades of red)

1044 when compared to the parental strains ΔaceE and ΔaceE Δpyc (shades of blue).

1045 Data represent the average of three biological replicates, error bars the standard

1046 deviation.

54

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

1047 - 1048 Figure 4: Increased availability of HCO3 improved anaplerotic flux via PEPCx in

1049 pyc-deficient C. glutamicum strains. Growth curves shown are based on the

1050 backscatter measurements in a microtiter cultivation system, inoculated at an OD600

1051 of 1. Symbols represent the backscatter means of biological triplicates (n=3). (A) 55

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

1052 Strain ΔaceE Δpyc Δppc was inoculated in CGXII medium containing 154 mM

- 1053 acetate and 222 mM glucose (shown in dark red) and, if indicated, 100 mM KHCO3

1054 was added (shown in light red). (B) The strains ΔaceE and ΔaceE Δpyc as well as

1055 the WT were cultivated in CGXII medium containing 154 mM acetate and 222 mM

1056 glucose adjusted to a pH of 6 (left) or pH of 8 (right) in comparison to cultivation at pH

1057 7. (C) The strains ΔaceE and ΔaceE ureD-E188* Δpyc were cultivated in CGXII

1058 media containing 154 mM acetate (left) or 154 mM acetate and 222 mM glucose

- 1059 (right) with and without the addition of KHCO3 . (D) The strains ΔaceE, ΔaceE Δpyc

1060 (left), ΔaceE ureD-E188* Δpyc (right) were inoculated in different growth media

1061 containing 154 mM acetate and 222 mM glucose (shades of blue), 100 mM gluconate

1062 (red) or 100 mM ribose (light red).

1063

56

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

1064 - 1065 Figure 5: The intracellular CO2/HCO3 pool significantly affects L-valine

1066 production. Investigations on the influence of carbonate on growth and L-valine

1067 production of C. glutamicum ΔaceE and ΔaceE ureD-E188* are shown. L-valine

1068 (filled bars) and L-alanine (dashed bars) titers in the supernatant after 28 h of

1069 cultivation in the microtiter cultivation system. Data represent average values from

1070 three independent biological replicates using ultra-high performance liquid

1071 chromatography (uHPLC).

1072

57

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

1073 1074 Figure 6: Refueling the TCA cycle to complement glucose intolerance of ΔaceE

1075 Δpyc. Growth curves shown are based on the backscatter measurements in a

1076 microtiter cultivation system, while symbols represent the backscatter means of

1077 biological triplicates (n=3). Strains ∆aceE and ∆aceE ∆pyc were inoculated to an

1078 OD600 of 1 in CGXII medium with 154 mM acetate and 222 mM glucose. (A) The TCA

1079 fueling carbon sources glutamate (3 mM Glu-Ala dipeptide), citrate (50 mM),

58

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

1080 succinate (100 mM) or malate (100 mM) were added to the medium to analyze their

1081 effect on lag phase complementation (shades of red). As a control, growth on 154

1082 mM acetate and 222 mM glucose is shown for ΔaceE (dark blue) and ΔaceE Δpyc

1083 (light blue) in each experiment. (B) Strains ∆aceE and ∆aceE ∆pyc were transformed

1084 with pJC1-aceA-linker-aceB or pJC1-aceB-linker-aceA plasmid for simultaneous

1085 overexpression of the glyoxylate shunt enzymes isocitrate lyase (aceA) and malate

1086 synthase (aceB), while transformation of ΔaceE Δpyc and Δpyc with pJC1-empty

1087 vector served as control. Cultures were inoculated to an OD600 of 1 in CGXII

1088 containing 154 mM acetate, 222 mM glucose and 25 μg/ml kanamycin.

59

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

1089 1090 Figure 7: Heterogeneous adaptation of C. glutamicum ΔaceE and ΔaceE Δpyc

1091 during growth on glucose and acetate. To identify a potential growing

1092 subpopulation the fluorescent dye PKH67 was used for membrane staining of ∆aceE

1093 and ∆aceE Δpyc prior to cultivation. Cells were cultivated in a microtiter cultivation

1094 system in the presence of 154 mM acetate and 222 mM glucose. As non-growing

1095 control, ∆aceE Δpyc was additionally incubated in CGXII containing solely 222 mM

1096 glucose. Cell growth (A) was online monitored using a microtiter cultivation system

60

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

1097 and staining intensities were measured on single cell level during cultivation using

1098 flow cytometry (n=3 biological replicates). (B) Shown is the frequency of PKH67

1099 staining intensities distributions of one representative culture of each sample over the

1100 time course of cultivation (A).

1101

61

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

1102 1103 Figure 8: Adaptive laboratory evolution of C. glutamicum ΔaceE Δpyc. Growth

1104 analysis of single inoculation steps obtained from the adaptive laboratory evolution

1105 (ALE) approach, which was performed with the strain ∆aceE ∆pyc in the presence of

1106 glucose and acetate. Growth curves are shown based on the backscatter

1107 measurements in a microtiter cultivation system. Symbols represent the backscatter

1108 means. (A) For growth analysis on the population level, glycerol stocks which were

1109 prepared during the ALE experiment, were used for the inoculation of a first pre-

1110 culture in BHI supplemented with 51 mM acetate. The second pre-culture in CGXII

1111 containing 154 mM acetate was then used for inoculation of the main culture in CGXII

1112 containing 154 mM acetate and 222 mM glucose (start OD600 of 1). Shown is the

1113 development of the relative lag phase duration after repetitive inoculations of ΔaceE

1114 Δpyc in media containing glucose and acetate. The difference between first doubling

1115 time of ΔaceE and ΔaceE Δpyc was set to 100%. For each inoculation, the difference

1116 of the lag phase compared to ΔaceE was calculated and is given in percent. (B)

1117 Single clones were isolated from batch number 7 (clone 1, black), 8 (clone 2, yellow)

1118 and 10 (clone 3, green) and were cultivated as mentioned above. ∆aceE (dark blue)

62

bioRxiv preprint doi: https://doi.org/10.1101/663856; this version posted June 9, 2019. The copyright holder for this preprint (which was not certified by peer review) is the author/funder. All rights reserved. No reuse allowed without permission.

1119 and ∆aceE ∆pyc (light blue) served as control. These single clones were further

1120 analyzed by genome sequencing (Table 7).

63