Anda di halaman 1dari 10

ARTICLE IN PRESS

Metabolic Engineering 11 (2009) 243–252

Contents lists available at ScienceDirect

Metabolic Engineering
journal homepage: www.elsevier.com/locate/ymben

Regular Article

Transcriptome analysis guided metabolic engineering of Bacillus subtilis


for riboflavin production
Shuobo Shi a,b,1, Tao Chen a,b,1, Zhigang Zhang c,d, Xun Chen a,b, Xueming Zhao a,b,!
a
Department of Biochemical Engineering, School of Chemical Engineering and Technology, Tianjin University, Tianjin 300072, China
b
Key Laboratory of Systems Bioengineering, Ministry of Education, Tianjin University, Tianjin 300072, China
c
BioTechnology Institute, 1479 Gortner Avenue, University of Minnesota, St. Paul, MN 55108, USA
d
Department of Chemical Engineering and Materials Science, 421 Washington Avenue SE, University of Minnesota, Minneapolis, MN 55455, USA

a r t i c l e in fo abstract

Article history: A comparative transcriptome profiling between a riboflavin-producing Bacillus subtilis strain RH33 and
Received 24 December 2008 the wild-type strain B. subtilis 168 was performed, complemented with metabolite pool and nucleotide
Received in revised form sequence analysis, to rationally identify new targets for improving riboflavin production. The pur operon
20 April 2009
(purEKBCSQLFMNHD) together with other PurR-regulated genes (glyA, guaC, pbuG, xpt-pbuX, yqhZ-folD,
Accepted 5 May 2009
Available online 13 May 2009
and pbuO) was all down-regulated in RH33, which consequently limited the supply of the riboflavin
precursors. As 5-phospho-ribosyl-1(a)-pyrophosphate (PRPP) strongly inhibits the binding of PurR to its
Keywords: targets, it was inferred that the reduced expression of PurR-regulated genes might be caused by a low
Riboflavin PRPP pool, which was subsequently confirmed by metabolite analysis. Thus, we selected and co-
Bacillus subtilis
overexpressed prs and ywlF genes in RH33, which are involved in the biosynthetic pathway of PRPP from
Transcriptome
ribulose-5-phosphate. This co-amplification led to an elevated PRPP pool and thus the increased
PRPP pool
Metabolic engineering transcript abundances of PurR-regulated genes participated in riboflavin precursor biosynthesis. The
riboflavin titer was increased by 25% (up to 15 g l!1) in fed-batch fermentation.
& 2009 Elsevier Inc. All rights reserved.

1. Introduction subtilis, currently the most competitive riboflavin producer, has


been widely adopted in the commercial riboflavin production
In the biotechnology industry, classical strain improvement processes (Knorr et al., 2007; Wu et al., 2007; Zamboni et al.,
has achieved a long history of success (Crueger and Crueger, 1984; 2003). The biosynthesis of riboflavin in B. subtilis occurs through
Peberdy, 1985), which relied on random mutagenesis and seven enzymatic steps starting from GTP and ribulose-5-phos-
selection. These methods are still very useful, especially with phate, which is shown in Fig. 1. The riboflavin producer B. subtilis
the development of efficient strain selection methods (Gall et al., was initially developed using the ‘‘classical’’ strain development
2008). However, unwanted changes in physiology and growth approach that relied on iterative cycles of random mutagenesis
retardation may occur alongside the desired improvements. Since and selection to create genetic diversity and identify improved
the introduction of metabolic engineering, a more rational riboflavin mutants (Perkins et al., 1991; Stahmann et al., 2000).
improvement approach emerges for microbial development Then, a number of conceivable strategies were carried out to
(Bailey, 1991; Stephanopoulos et al., 1998). construct high-level riboflavin-producing B. subtilis strains. These
Riboflavin (vitamin B2) serves as a precursor for the synthesis include enhancement of both gene dosages and transcriptional
of the coenzymes flavin mononucleotide (FMN) and flavin adenine level of riboflavin operon in the mutants (Perkins et al., 1999),
dinucleotide (FAD), which are used as electron acceptors for many constitutive expression of the key gene (ribA) in riboflavin
oxidoreductases (Stahmann et al., 2000). As such, riboflavin is biosynthetic pathway (Humbelin
+ et al., 1999), enhancing energy
required for a wide variety of cellular processes and is supple- generation and reducing maintenance metabolism (Zamboni
mented for feed and food fortification purposes in humans and et al., 2003), increasing precursor supply by modulating carbon
animals to maintain health. The Gram-positive bacterium Bacillus flow through pentose phosphate pathway (Zamboni et al., 2004a;
Zhu et al., 2006), and deregulating gapB expression by ccpN
knockout based on screening B. subtilis transposon mutants
! Corresponding author at: Department of Biochemical Engineering, School of
(Tännler et al., 2008).
Chemical Engineering and Technology, Tianjin University, Tianjin 300072, China.
However, these strategies were not based on a comprehensive
Fax: +86 22 27406770.
E-mail address: xmzhao@tju.edu.cn (X. Zhao). analysis of the complex microbial metabolism and regulation,
1
These authors contributed equally to this work. which would further facilitate the successful and efficient

1096-7176/$ - see front matter & 2009 Elsevier Inc. All rights reserved.
doi:10.1016/j.ymben.2009.05.002
ARTICLE IN PRESS
244 S. Shi et al. / Metabolic Engineering 11 (2009) 243–252

Fig. 1. Schematic overview of expression profiles of genes involved in relevant pathways of riboflavin production. The numbers are the ratios of the comparative expression
levels in B. subtilis strain RH33 vs. 168. Blue lettering indicates downregulation, red indicates up-regulation, and black indicates without notable changes. The genes that
were underlined were selected for overexpression. G6P, glucose-6-phosphate; F6P, fructose-6-phosphate; GAP, D-glyceraldehyde 3-phosphate; PGA, 3-phosphoglycerate;
PYR, pyruvic acid; AcoA, acetyl coenzyme A; Ser, serine; Gly, glycine; 10-formyl-THF, 10-formyl tetrahydrofolate; gluconate-6-P, 6-phospho-D-gluconate; Ru-5-P, ribulose-5-
phosphate; Ribo-5-P, ribose-5-phosphate; PRPP, phosphoribosylpyrophosphate; Gln, glutamine; X5P, xylulose 5-phosphate; E4P, D-erythrose 4-phosphate; IMP, inosine
monophosphate; XMP, xanthosine monophosphate; GMP, guanosine mono-phosphate; GTP, guanosine tri-phosphate; DARPP, 2,5-diamino-6-ribosylamino-4(3H)-
pyrimidinone-50 -phosphate; ARPP, 5-amino-6-(50 -phosphoribosylamino)uracil; ArPP, 5-amino-6-(50 -phosphoribitylamino)uracil; ArP, 4-(1-D-ribitylamino)-5-amino-2,6-
dihydroxypyrimidine; DHPB, 3,4-dihydroxy-2-butanone 4-phosphate; DRL, 6,7-dimethyl-8-ribityl-lumazine; FAD, flavin adenine dinucleotide (For interpretation of the
references to colour in this figure legend, the reader is referred to the web version of this article.).

microbial metabolic engineering (Nielsen and Olsson, 2002). Table 1


Recently, as a holistic and discovery-driven approach, transcrip- Strains and plasmids used in this study.
tome analysis has been successfully applied in metabolic
Strain or Description of genotype Source
engineering for rationally probing the complex gene and meta- plasmid
bolic regulatory networks. Novel target genes have been identified
to optimize microbial production strains (Choi et al., 2003; Harris Strains
et al., 2009; Izallalen et al., 2008; Jaluria et al., 2007; Park et al., B. subtilis Wild-type BGSC
168
2007; Peebles et al., 2009; Sindelar and Wendisch, 2007; Wierckx B. subtilis Emr, Cmr, containing multiple riboflavin operons Laboratory
et al., 2008). RH33 stock
Here we took a strategy to increase riboflavin production in a B. subtilis Emr, Cmr, Spcr, containing a P43-prs (CDS) fragment This study
riboflavin-producing B. subtilis strain based on comparative RH33-Prs integrated in the chromosome of RH33
B. subtilis Emr, Cmr, Spcr, containing a P43-prs-ywlF (CDS) This study
transcriptome analysis between a riboflavin high-producer RH33
RH33-PY fragment integrated in the chromosome of RH33
and wild type 168, integrated with DNA sequencing and E. coli Top10 Host strain for constructing plasmids Laboratory
metabolite pool measurement. Our integrated approach allowed stock
us to understand genome-wide transcriptional differences under- Plasmids
lying strain performance, and to identify potential metabolic pUC18 Ampr BGSC
bottlenecks for riboflavin production. We rationally overexpressed pSG1192 Ampr, Spcr BGSC
two genes simultaneously to activate precursor purine biosynth- pHPL10 pHP13, containing a P43 promoter Laboratory
stock
esis pathways by modulating global regulator activity via
pRPU10 Ampr, containing a prs (CDS) fragment This study
metabolite pool manipulation. This strategy was capable of pRPU12 Ampr, containing a P43-prs (CDS) fragment This study
elevating riboflavin titer by 25% in B. subtilis RH33 (up to 15 g l!1) pRPU13 Ampr, Spcr, containing a P43-prs (CDS) fragment This study
in fed-batch fermentation. pRPU14 Ampr, containing a P43-prs-ywlF (CDS) fragment This study
pRPU15 Ampr, r
Spc , containing a P43-prs-ywlF (CDS) fragment This study

BGSC, Bacillus Genetic Stock Center (http://www.bgsc.org/).


2. Materials and methods

2.1. Strains and plasmids


pDG364 (BGSC, http://www.bgsc.org/). Then the vector pRB63
Bacterial strains and plasmids used in this work are listed in was integrated into the B. subtilis RH13 chromosome at the native
Table 1. Two different riboflavin-producing mutants B. subtilis rib operon locus by single crossover and selected at 5 mg ml!1
RH13 and B. subtilis J617 were derived from B. subtilis 168 by chloramphenicol. Subsequently, the copy number of the
multiple rounds of selection with azaguanine (Azr), decoyinine integrated pRB63 was increased by selecting colonies that grew
(Dcr) and roseoflavin (RoFr) for resistance mutations that at higher chloramphenicol concentrations. Colonies that were able
deregulate the riboflavin biosynthetic pathway. An integration to grow up with 40 mg ml!1 chloramphenicol were isolated, and
vector pRB63 was constructed by inserting a native rib operon into one was denoted as B. subtilis RH13H[pRB63]n. Another
ARTICLE IN PRESS
S. Shi et al. / Metabolic Engineering 11 (2009) 243–252 245

integration vector pRB62 carried erythromycin resistant was 0.03 g l!1 ZnCl2 " 7H2O, 0.04 g l!1 MnCl2 " 4H2O. The cells were
integrated into the B. subtilis J617 chromosome at the native grown in the batch mode for 6 h after inoculation, and then the
thrC locus by double crossover. The vector was constructed by feeding process was continued from 6 to 48 h. The feed solution
insertion of a 4.4-kb P43-modified rib operon (substituting the contained 650 g l!1 glucose, 10 g l!1 yeast extract, 10 g l!1 NaNO3,
native ribP1 promoter and RFN regulatory element with the 5 g l!1 NH4NO3, 5 g l!1 KH2PO4, 5 g l!1 K2HPO4, and 2 g l!1
constitutive P43 promoter) into pDG1664 (BGSC, http:// MgSO4 " 7H2O. The residual glucose concentration in the fermen-
www.bgsc.org/). The resulting strain was denoted as B. subtilis tor was maintained no more than 10 g l!1 by controlling the
J617H[pRB62]. Subsequently, B. subtilis RH13H[pRB63]n and B. glucose feed rate manually. The cultivation was carried out for
subtilis J617H[pRB62] were fused by protoplast fusion. The fusants 48 h at the agitation rate of 900 rpm, with an aeration rate of
were selected in solid regeneration medium containing 1.1 vvm and temperature of 41 1C. The head pressure is maintained
40 mg ml!1 chloramphenicol and 5 mg ml!1 erythromycin. One at 0.5 atm. pH was maintained at 6.8 with 1 M H2SO4 and 10%
clone was denoted as B. subtilis RH33, which was used as the ammonia. All the experiments were carried out independently in
parental strain in our study. biological triplicates, and the reported results were the average of
Plasmids were constructed by the following procedures. Firstly, three replicate experiments.
polymerase chain reaction (PCR) was used to amplify structural
genes of prs and ywlF with the following primers: prs-up 50 - 2.3. Transcriptome analysis
GGGGCCCGGGCCAGAGCGAGACAAGTAAA, prs-down 50 -GGG-
GGAGCTCGCTAGCTCCTATTACAAACAATACCCA, ywlF-up 50 -GGG- About 25 ml of the culture was used for preparation of total
GCCCGGGGCTAGCGGCTGCGCGGTCAATA, ywlF-down 50 -GGGG- RNA, which was extracted using RiboPureTM-Bacteria Kit (Ambion,
GAGCTCGCGGCCGCTTGTTTCAATTCCGCTTGGTC, based on the UK). Total RNA was stored at !70 1C until use. The quantity and
published B. subtilis 168 genome sequence (Kunst et al., 1997). quality of RNA was analyzed by UV spectrophotometry and
The prs PCR product was ligated into pUC18 using XmaI and SacI denaturing formaldehyde agarose gel electrophoresis, respec-
restriction sites to construct pRPU10. The constitutively expressed tively. Probe sets on the B. subtilis Genome Array (Antisense)
P43 promoter (Wang and Doi, 1984) was obtained from the vector were designed based on the wild-type B. subtilis 168 genome
pHPL10 (unpublished work) after digested with BamHI and XmaI. sequence data of Kunst et al. (1997) by Affymetrix (Santa Clara,
It was then ligated in the BamHI and XmaI restriction sites of CA, USA). An aliquot of 10 mg B. subtilis total RNA was used to
pRPU10 to get plasmid pRPU12. The ywlF PCR product was synthesize first-strand cDNA with random primers and Super-
digested with NheI and SacI, and cloned into NheI-SacI-sites of Script II reverse transcriptase. Then the cDNA was fragmented to
pRPU12 to construct pRPU14. To facilitate further research, a 50–200 bp and labeled by biotin. After hybridization at 45 1C for
spectinomycin resistance gene (spc) from pSG1192 was inserted at 16 h at 60 rpm, the microarray slides were washed and stained on
the SalI site of pRPU12 and pRPU14 to give plasmids pRPU13 and Affymetrix Fluidics Station 450. The scanned images were
pRPU15, respectively. All DNA manipulations were carried out as obtained with the GeneChip Scanner 3000 (Affymetrix) and were
described previously (Sambrook and Russell, 2001). LB medium analyzed using the default setting of GeneChip Operating Soft-
supplemented with corresponding antibiotics was used as the ware (GCOS 1.4). Then a LOWESS normalization method was
standard medium for plasmid construction in Escherichia coli. The performed to normalize the different arrays using dChip software.
plasmids were transformed into competent cells of B. subtilis The differentially expressed genes were identified through over-
RH33 as described previously (Anagnostopoulos and Spizizen, lapping gene analysis of biological duplicate experiments using a
1961). 2-fold change as an empirical criterion.

2.2. Growth conditions 2.4. Quantitative RT-PCR

For transcriptome and quantitative RT-PCR analyses, 5-phos- To validate transcriptome results, relative abundances of
pho-ribosyl-1(a)-pyrophosphate (PRPP) pool, and riboflavin mea- selective transcripts were measured by quantitative reverse
surements, a preculture of B. subtilis RH33 and B. subtilis 168 were transcription-PCR, which was carried out by an iCycler (Bio-Rad,
inoculated from a fresh LB agar plate and cultured to exponential USA) using the QuantiTect SYBR Green RT-PCR kit (Qiagen,
growth period in LBG medium (LB medium with 1% glucose) at Germany) according to the manufacturer’s instructions. In brief,
37 1C. The intracellular metabolites, PRPP, ribose-5-phosphate and 100 ng of DNA-free total RNA was used in a total reaction volume
ribulose-5-phosphate, were determined during exponential of 50 ml with 0.4 mM of each primer (Supplementary Table 1). The
growth period in minimal medium with 20 g l!1 glucose, 2 g l!1 fold change of each transcript in each sample relative to the
(NH4)2SO4, 13.1 g l!1 KH2PO4, 6 g l!1 K2HPO4, 1.2 g l!1 NaC6H5O7 " control sample was measured in triplicates, normalized to internal
2H2O, 10 mg l!1 MgSO4 " 7H2O and supplemented with trypto- control gene gapA and calculated according to the comparative Ct
phan, phenylalanine and tyrosine (25 mg l!1 each). method (Livak and Schmittgen, 2001).
Fed-batch fermentation was performed in B. subtilis strains for
riboflavin overproduction. The strain was revived by growing on 2.5. Metabolite concentration measurement
LB agar slants. The seed cultures of the revived B. subtilis strains
were prepared in 500-ml shake flasks at 41 1C with 240 rpm. The Immediately upon harvest, 5 ml of cell sample was added to
seed medium contained 20 g l!1 glucose, 5 g l!1 yeast extract, 15 ml of quenching fluid containing 70 mM HEPES in 60% (v/v)
10 g l!1 NaNO3, 5 g l!1 NH4NO3, 1 g l!1 KH2PO4, 1.65 g l!1 K2HPO4, aqueous methanol. The samples were centrifuged to separate the
1.5 g l!1 MgSO4 " 7H2O, 0.03 g l!1 FeCl2, 0.04 g l!1 MnCl2 " 4H2O, quenching fluid and the remaining cell pellets were resuspended in
0.04 g l!1 ZnCl2 " 7H2O. After 14 h incubation at 41 1C, 150 ml of the 35% (v/v) perchloric acid. After one freeze–thaw cycle the sample
seed culture was transferred into in a 5-l bioreactor (Bao Xing, was neutralized by 5 M K2CO3. The precipitate was removed by
China) containing 2350 ml fermentation medium. The initial another centrifugation, and resulting supernatants were stored at
medium of fed-batch fermentations contained 20 g l!1 glucose, !80 1C until analysis. Measurement of intracellular PRPP concen-
5 g l!1 yeast extract, 10 g l!1 NaNO3, 5 g l!1 NH4NO3, 4 g l!1 trations was carried out using the methods described by Kornberg
KH2PO4, 7.5 g l!1 K2HPO4, 1.5 g l!1 MgSO4 " 7H2O, 0.03 g l!1 FeCl2, et al. (1955). The ribose-5-phosphate and ribulose-5-phosphate
ARTICLE IN PRESS
246 S. Shi et al. / Metabolic Engineering 11 (2009) 243–252

were measured using a Luminescence Spectrophotometer (Hoque Table 2


et al., 2005). Validation of microarray data by quantitative RT-PCR.

Genes Quantitative RT-PCR Microarray


2.6. Enzyme activity assay in crude extracts
purF 0.2570.01 0.3870.01
purD 0.1170.01 0.5070.01
The crude cell extract was prepared as described previously glyA 0.4270.02 0.6870.03
(Fisher and Mangasanik, 1984). For enzyme assay, strains were folD 0.4470.02 0.6070.04
cultivated at 37 1C in minimal medium. Quantification of PRPP prs 0.9570.02 1.0070.03
synthetase activity, ribose-5-phosphate isomerase activity, and ywlF 0.8770.01 0.9070.06
ribA 15970.37 13.2770.08
transketolase activity were performed as described (Hove-Jensen
narG 2.9970.05 6.4870.06
and Maigaard, 1993; Iida et al., 1993; Sakuma et al., 1991). Specific
activity was expressed as nmol min!1 mg!1 protein. Protein Fold change of each transcript in B. subtilis RH33 relative to 168 was reported.
content was determined according to the Bradford method.

regulated to 13.27 and 6.48-fold, respectively. The genes prs and


2.7. Analytical methods
ywlF were identified not to be differentially expressed (1.00 and
0.90) by microarray, which was confirmed by quantitative RT-PCR
Cell growth (OD600 nm) was monitored with a spectrophot-
(0.95 and 0.87). Furthermore, by quantitative RT-PCR, the mean
ometer. Cell growth rate was calculated by log-linear regression
expression levels of purF, purD, glyA, and folD were identified to be
analysis of OD600 nm versus time. Glucose concentration was
0.25, 0.11, 0.42, and 0.44, respectively. All the four genes were
determined using a glucose analyzer (Model-SBA40, Shandong,
identified as significantly down-regulated (0.38, 0.50, 0.68, and
China). For riboflavin measurements, the samples were diluted
0.60) in B. subtilis RH33 by transcriptome analysis. Therefore, the
with 0.05 M NaOH to the linear range of the spectrophotometer,
correlation between quantitative RT-PCR and microarray was
and centrifuged at 12,000 rpm for 2 min to remove the cells. Then
good (Table 2), suggesting the validity of the microarray gene
the OD444 nm was immediately measured. The culture broth was
expression measurements.
centrifuged and the resulting supernatant was used for measure-
The differentially expressed genes fell into nearly all functional
ment of the concentrations of extracellular metabolites (acetate,
categories (Supplementary Table 2). The largest group with
acetoin, etc.) by using a HPLC system (Agilent, HP1100, USA)
altered transcriptional levels in the B. subtilis RH33 was a group
equipped with a UV–visible light monitor (Bai et al., 2004).
of 83 genes encoding for the transport/binding proteins and
lipoproteins, all of which are associated with the cell membrane.
2.8. Nucleotide sequencing Forty-eight affected genes belonged to the group involved in
metabolism of amino acids and related molecules, especially
Selected genes involved in riboflavin biosynthesis and cellular aspartate, cysteine, glutamate, histidine, leucine, threonine,
central metabolism were sequenced for both regulatory and tyrosine, and serine biosynthesis genes. Other large functional
coding regions, using the traditional Sanger method by Beijing groups changed significantly included 24 genes associated with
AuGCT biotechnology Co., Ltd., China. motility, 23 genes associated with carbohydrate metabolism
(notably myoinositol and acetoin metabolism), 23 genes asso-
ciated with metabolism of nucleotides and nucleosides (mainly
3. Results and discussion purine biosynthetic genes), 32 genes associated with transcription
regulation and 27 genes associated with phage-related function.
3.1. Comparative transcriptome analysis Therefore, it seems that complex transcriptional regulation
mechanism(s) underpinned the strain performance differences,
Transcriptome profiles of a riboflavin-producing B. subtilis RH33 directly or indirectly affecting riboflavin production. Some of
was compared to wild-type B. subtilis 168 on DNA microarray using these transcriptional effects will be discussed in detail as follows.
mid-exponential phase cell samples (OD600 nm#0.90) grown in LBG
medium. In B. subtilis RH33, this condition corresponded to the
state of high riboflavin production, which was coupled to cell 3.2. Nucleotide sequencing of selected genes
growth. The specific growth rate of RH33 and 168 were 1.5770.15
and 1.6170.15 h!1, respectively. During mid-exponential phase, the Since B. subtilis RH33 was developed mostly by repeated
specific growth rates were nearly constant, without discernible random mutagenesis and selection, mutations that affect gene
accumulation of byproducts such as acetate (data not shown). Our function were also expected to occur. The identification of
past experiences showed that there was no significant difference in mutational changes is necessary to fully understand riboflavin
transcriptional profiles at different time-points of mid-exponential synthesis processes in RH33. However, the unknown random
phase. mutations may affect the activity of gene product, but may not
The microarray experiment was very reproducible. The Pearson affect gene expression per se and thus such mutations may not be
correlation coefficient between the biological duplicates was detected by microarray. To identify possible beneficial mutations
about 0.95; the overlapping of differentially expressed genes for riboflavin production in RH33, we selected 67 genes for DNA
identified using an empirical 2-fold criterion from each experi- sequencing, including both regulatory and coding regions. These
ment was more than 90%. It was found that 619 genes showed genes participated in riboflavin biosynthesis, precursor supply
significant variations at the genome-wide transcriptional level and cell central metabolism, which were simply grouped into six
between RH33 and 168, representing about 15% of B. subtilis functional categories (Table 3): glycolysis and TCA cycle, pentose
genome. Quantitative RT-PCR experiment was performed in phosphate pathway, purine pathway and other PurR-regulated
parallel to validate microarray data (Table 2). The average fold genes, riboflavin biosynthesis and transport, glycine biosynthesis,
change of ribA and narG transcripts, were identified to be 159 and and nitrogen metabolism. The nucleotide sequences of six genes
2.99 in RH33 relative to 168, respectively. This was consistent were identified to contain mutations in RH33, including ribC (the
with microarray data, since both genes were significantly up- riboflavin kinase and FAD synthase gene), tkt (the transketolase
ARTICLE IN PRESS
S. Shi et al. / Metabolic Engineering 11 (2009) 243–252 247

Table 3
Relative expression changes and nucleotide sequence mutations of selected genes directly or indirectly participated in riboflavin biosynthesis in RH33.

Gene DFold Gene product Mutations

Glycolysis and TCA cycle


pycA 1.32 Pyruvate carboxylase No
pdhA 0.68 Pyruvate dehydrogenase (E1 alpha subunit) No
pdhB 0.72 Pyruvate dehydrogenase (E1 beta subunit) No
pdhC 0.88 Pyruvate dehydrogenase (dihydrolipoamide acetyltransferase E2 subunit) No
pdhD 0.88 Pyruvate dehydrogenase/2-oxoglutarate dehydrogenase (dihydrolipoamide dehydrogenase E3 subunit) No
gapB 1.08 Glyceraldehyde-3-phosphate dehydrogenase No
citZ 1.96 Citrate synthase II (major) No
citG 0.47 Fumarate hydratase No
pta 0.77 Phosphotransacetylase No
ldh (lctE) 1.17 L-lactate dehydrogenase Gly29Gly
ackA 0.80 Acetate kinase No
alsS 0.82 Alpha-acetolactate synthase No
alsD 0.76 Alpha-acetolactate decarboxylase No
ptsG 1.07 PTS glucose-specific enzyme IICBA component No

Pentose phosphate pathway


zwf 1.44 Glucose-6-phosphate 1-dehydrogenase No
gntZ 0.29 6-Phosphogluconate dehydrogenase No
yqjI 1.11 Similar to 6-phosphogluconate dehydrogenase No
tkt 0.89 Transketolase 17delA
rpe 0.90 Ribulose-5-phosphate 3-epimerase Val4Asp
ywlF 0.90 Ribose-5-phosphate isomerase B No
prs 1.00 Phosphoribosylpyrophosphate synthetase No

Nitrogen metabolism
yweB (rocG) 4.82 Glutamate dehydrogenase No
glnA 0.81 Glutamine synthetase Glu65Lys
gltC 0.82 Transcriptional activator of the glutamate synthase operon No
gltA 0.46 Glutamate synthase (large subunit) (EC: 1.4.1.13) No
gltB 0.37 Glutamate synthase (small subunit) (EC: 1.4.1.13) No
nasF 1.62 Uroporphyrin-III C-methyltransferase No
nasE 1.59 Assimilatory nitrite reductase (subunit) No
nasD 1.34 Assimilatory nitrite reductase (subunit) No
nasC 1.31 Assimilatory nitrate reductase (catalytic subunit) No
nasB 1.81 Assimilatory nitrate reductase (electron transfer subunit) No
narG 6.48 Nitrate reductase (alpha subunit) No
narH 13.59 Nitrate reductase (beta subunit) No
narJ 16.35 Nitrate reductase (protein J) No
narI 19.27 Nitrate reductase (gamma subunit) No

Glycine biosynthesis
serA 0.37 Phosphoglycerate dehydrogenase No
serC 0.71 Phosphoserine aminotransferase No
glyA 0.68 Serine hydroxymethyltransferase No
folD 0.6 Methylenetetrahydrofolate dehydrogenase/methenyltetrahydrofolate cyclohydrolase No

Purine biosynthesis pathway and other PurR-regulated genes


purE 0.47 Phosphoribosylaminoimidazole carboxylase I (EC: 4.1.1.21) No
purK 0.48 Phosphoribosylaminoimidazole carboxylase II (EC: 4.1.1.21) No
purB 0.28 Adenylosuccinate lyase (EC: 4.3.2.2) No
purC 0.21 Phosphoribosylaminoimidazole succinocarboxamide synthetase (EC: 6.3.2.6) No
yexA (purS) 0.23 Required for phosphoribosylformylglycinamidine synthetase activity No
purL 0.24 Phosphoribosylformylglycinamidine synthetase II (EC: 6.3.5.3) No
purQ 0.24 Phosphoribosylformylglycinamidine synthetase I (EC: 6.3.5.3) No
purF 0.38 Glutamine phosphoribosylpyrophosphate amidotransferase (EC: 2.4.2.14) No
purM 0.30 Phosphoribosylaminoimidazole synthetase (EC: 6.3.3.1) No
purN 0.37 Phosphoribosylglycinamide formyltransferase (EC: 2.1.2.2) No
purH 0.50 Phosphoribosylaminoimidazole carboxy formyl formyltransferase/inosine-monophosphate cyclohydrolase No
purD 0.50 Phosphoribosylglycinamide synthetase No
purR 0.98 Transcriptional repressor of the purine operon No
purA 0.55 Adenylosuccinate synthetase No
guaB 0.88 Inosine-monophosphate dehydrogenase No
guaC (yumD) 0.23 GMP reductase Leu305Ile
yebB (pbuG) 0.42 Hypoxanthine/guanine permease No
Xpt 0.36 Xanthine phosphoribosyltransferase No
pbuX 0.3 Xanthine permease No
yqhZ 0.74 Probable transcription termination No
ytiP (pbuO) 0.24 Unknown No

Riboflavin biosynthesis and transport


ribA 13.27 GTP cyclohydrolase II and 3,4-dihydroxy-2-butanone 4-phosphate synthase (EC: 3.5.4.25) No
ribB 13.19 Riboflavin synthase alpha chain [EC: 2.5.1.9] No
ribC 3.26 Riboflavin kinase and FAD synthase (EC: 2.7.1.26 and EC: 2.7.7.2) Gly199Asp
ribG 21.74 Diaminohydroxyphosphoribosylaminopyrimidine deaminase/5-amino-6-(5-phosphoribosylamino)uracil No
reductase [EC: 3.5.4.26 1.1.1.193]
ribH 10.62 Riboflavin synthase (beta subunit) (EC: 2.5.1.9) No
ribT (ribD) 7.13 Reductase No
ypaA 4.70 A possible transporter of riboflavin No
ARTICLE IN PRESS
248 S. Shi et al. / Metabolic Engineering 11 (2009) 243–252

gene), guaC (GMP reductase), ldh (L-lactate dehydrogenase), glnA 1, glucose-6-phosphate enters the pentose phosphate pathway
(the glutamine synthetase gene), and rpe (the ribulose-5- after being oxidated to gluconate-6-phosphate by glucose-6-
phosphate 3-epimerase gene). The mutations in the ribC and tkt phosphate dehydrogenase (zwf) and then is subsequently oxidized
were considered as beneficial mutations for riboflavin to ribulose-5-phosphate by gluconate-6-phosphate dehydrogen-
biosynthesis as discussed below, while causal links to riboflavin ase. It was found that the transcription of gntZ, which was
production were not obvious for the other mutations except the considered encoding the major 6-P-gluconate dehydrogenase, was
synonymous mutation in gene ldh, Gly(GGG)-29-Gly(GGA). significantly down-regulate (0.29-fold) in B. subtilis RH33. How-
ever, according to a recent report (Zamboni et al., 2004b), its
isoenzyme gene yqjI was shown to have the major 6-P-gluconate
3.3. Transcriptional regulation and nucleotide mutation contributed dehydrogenase activity in B. subtilis, which was moderately up-
to the riboflavin overproducing traits in B. subtilis RH33 regulated in RH33 (1.1-fold) to compensate the enzyme activity
and carbon flux. It should be noted that the expression of the
First, in the riboflavin biosynthesis metabolism, the transcrip- transketolase gene tkt did not alter significantly in magnitude
tion of the rib genes, which are organized in an operon and encode (0.89-fold). However, there was a frame-shift mutation in its
the riboflavin biosynthesis pathway in B. subtilis, were strongly coding region. It had been shown that the B. subtilis RH33 grew
up-regulated in RH33 (Table 3; Fig. 1). The expression of the rib very slowly in minimal medium supplement with xylose or
operon genes were significantly up-regulated by 13.2 (ribA), 13.2 arabinose as the sole carbon source (unpublished data), which
(ribB), 10.6 (ribH), and 21.7-fold (ribG), reflecting their over- may be related to a dramatic lose of the transketolase activity.
expression and direct contribution to riboflavin overproducing Indeed, B. subtilis RH33 showed only 4% transketolase activity
phenotype. Since rib operon genes are directly involved in compared with that of B. subtilis 168 (177 nmol min!1 mg!1
riboflavin biosynthesis (Perkins et al., 1999), it is likely that the protein). The low activity of transketolase may decrease the shunt
up-regulation of rib operon genes positively contributed to the of pentose back into glycolysis, allowing cells to redirect more
riboflavin production. carbon flux through the oxidative pentose pathway to the purine
In B. subtilis, the rib operon has an untranslated regulatory nucleotide biosynthesis pathway, which provided the precursors
leader region of 300 base pairs in front of its first gene, ribG. The of riboflavin production (Kamada et al., 2001). In short, tkt
leader sequence, a conserved regulatory element called RFN contributed to riboflavin overproduction trait of RH33, not at the
element, is responsible for negative regulation of the operon by level of transcriptional regulation, but through frame-shift
FMN binding, which prevents the expression of the downstream mutation to reduce enzyme activity of gene tkt.
genes by transcriptional or translational mechanisms (Nudler and
Mironov, 2004). Revealed by DNA sequencing, there are no
mutations in the RFN element in RH33. The up-regulation of the 3.4. Identification of metabolic bottleneck for increasing riboflavin
bifunctional flavokinase/FAD synthase gene ribC (3.3-fold), which production in B. subtilis RH33
converts riboflavin to FMN and FAD, is not beneficial for high
riboflavin accumulation. However, DNA sequencing indicated that Purine nucleotides, involving in biosynthesis of riboflavin by
ribC of RH33 had a Gly-199-Asp mutation, which was coincident providing the immediate precursor GTP, are synthesized by purine
with the report of Bresler et al. (1973) and could lead up to 95% genes (purEKBCSQLFMNHD). Unexpectedly, the purine genes
drop in enzyme activity. Thus, the up-regulation of rib operon in together with glycine biosynthesis genes (serA, serC, glyA, folD)
RH33 may be partly attributed to diminished feedback inhibition (Saxild et al., 2001) were all greatly down-regulated in RH33
of FMN resulting from the low flavokinase activity of the mutated (Table 3; Fig. 1). The down-regulation of purF, purD, glyA, and folD
ribC. In addition to the riboflavin biosynthesis genes, the RFN were also revealed by quantitative RT-PCR experiment (Table 2).
element was also observed in the upstream of ypaA (Nudler and Thus, the down-regulated purine and glycine (another building
Mironov, 2004) which was predicted to encode a transporter of block for riboflavin) biosynthesis genes would limit the supply of
riboflavin. In RH33, ypaA gene was up-regulated by 4.7-fold. Based the precursors for riboflavin overproduction (Heefner et al., 1988;
on these data, it was likely that the FMN concentration was Jimenez et al., 2005; Mateos et al., 2006; Schlupen et al., 2003;
maintained at a low level, reflected from the highly up-regulation Stahmann et al., 2000).
of rib operon, and ypaA gene that all have the RFN element (Lee Nonetheless, it was interesting to note that the down-
et al., 2001; Vogl et al., 2007). The mutation in ribC, which leads regulated purine, glycine, and other genes involved in purine
up to 95% drop in enzyme activity (Bresler et al., 1973), may metabolism (guaC, pbuG, xpt-pbuX, yqh, and pbuO) were all
directly contribute to the low FMN pool. negatively regulated by global regulator PurR in B. subtilis (Saxild
Second, transcriptome profiling revealed that the mRNA et al., 2001). The nucleotide sequencing results indicated that
expression level of the genes involved in the central carbon there were no mutations in the regulatory region (PRPP binding
metabolism had small variations between wild-type strain 168 motif) and coding region of the purR gene. Among all the PurR-
and the riboflavin overproducing strain RH33 (Table 3; Fig. 1). regulated genes only guaC had a point mutation (Table 3). It was
These genes may be highly expressed in both strains, as reported that the defective GMP reductase (guaC) that reduces
qualitatively reflected from their high absolute fluorescence conversion of GMP to IMP had been identified in a B. subtilis
intensities in one-color DNA microarray hybridization. Though riboflavin overproducer (Nygaard, 1993). However, the mutation
riboflavin was over-produced, few genes involved in central of guaC in B. subtilis RH33 was neutral (from Leu to Ile) and
metabolism showed differential expression in strain RH33. It probably had no effect on protein function.
may indicate that precursor supply due to the carbon flux through Since no mutations were found in the regulatory region and
central metabolism was sufficient in RH33 under riboflavin coding region of the PurR-regulated genes (including purR) except
overproduction condition. guaC, the down-regulation of those genes may be due to the
Alteration of carbon flow into the pentose phosphate pathway repression by PurR. It binds to the operator sites (PurBox) of its
will affect the riboflavin production because the pentose phos- target genes and inhibits their transcription. According to the
phate pathway is a major source for the supply of ribose current model, the PRPP can bind PurR to prevent its binding to
(or ribulose-5-phosphate), which is the starting point of riboflavin DNA and lead to derepression of PurR-regulated genes (Weng
biosynthesis (Sauer et al., 1996; Zhu et al., 2006). As shown in Fig. et al., 1995). Therefore, it can be deduced that the low expression
ARTICLE IN PRESS
S. Shi et al. / Metabolic Engineering 11 (2009) 243–252 249

of PurR-regulated genes might be caused by a small cellular PRPP The selected prs gene was overexpressed in B. subtilis RH33
pool. Subsequently, metabolite analysis was carried out and tested under the control of the P43 promoter to construct the strain
this hypothesis (Table 4). The PRPP pool in B. subtilis RH33 was B. subtilis RH33-Prs. The enzyme activities of PRPP synthetase
only 24% that of B. subtilis 168 measured at the conditions as the were 3.070.7 and 10.172.0 nmol min!1 mg!1 protein for B.
microarray experiments. In addition, PRPP itself is an important subtilis RH33 and B. subtilis RH33-Prs, respectively, indicating that
metabolite precursor for purine biosynthesis as it is required in PRPP synthetase was overexpressed successfully.
the de novo and salvage pathways of purine biosynthesis (Jimenez To assess the effects of the manipulation on the physiological
et al., 2008). It was proposed that enhancing PRPP pool would parameters, batch cultivations were carried out in minimal
derepress the expression of pur operon and glycine biosynthesis medium. The results displayed only minor variability in the
genes, thus increasing the precursors supply and leading to specific cell growth rate, specific glucose uptake rate, specific
increased production of riboflavin in B. subtilis RH33. At present, organic acids secretion rate, and specific riboflavin production rate
there is only one report employing the strategy of modulating the between B. subtilis RH33 and B. subtilis RH33-Prs (Table 6).
PRPP pool for riboflavin production (Jimenez et al., 2008) in a Unexpectedly, the PRPP concentration was almost constant in the
hypothesis-driven approach. mutant (Table 4). These indicated that overexpression of prs gene
alone had little effect on the supply of PRPP and on specific
riboflavin production rate.

3.5. Effect of prs overexpression on riboflavin production


3.6. Enhancement of riboflavin production by co-expression of prs
PRPP is an important metabolite required for purine, pyrimi- and ywlF genes
dine, tryptophan, and histidine biosynthesis. As discussed above,
the low concentration of PRPP may be the bottleneck for further As the strain B. subtilis RH33-Prs did not show the expected
increasing riboflavin production in B. subtilis RH33. The formation traits in improving PRPP supply, to diagnose the problem(s),
of PRPP is catalyzed by the enzyme PRPP synthetase, which is further measurements of precursor concentrations of the meta-
encoded by prs gene. The expression level of prs was not notably bolites were carried out to elucidate the limitations. Compared to
altered. In addition, the enzyme activity assay showed that the RH33, ribulose-5-phosphate had similar levels but ribose-
enzyme activity of RPPP synthetase was almost the same between 5-phosphate had significantly lower concentrations in RH33-
B. subtilis RH33 and B. subtilis 168 (Table 5). Therefore, it was Prs (Table 4). The insufficient ribose-5-phosphate pool was
reasoned that amplification of prs gene was expected to increase identified as the limiting factor for the indistinctive effect on
the supply of PRPP and consequently result in increased riboflavin increasing RPPP concentration and riboflavin production in B.
production. subtilis RH33-Prs.

Table 4
Intracellular metabolite concentration in cell extracts of B. subtilis.

Metabolite [nmol mg!1 (dry wt)] 168 RH33 RH33-Prs RH33-PY

PRPP 2.5070.36a 0.8870.31a 0.9170.32a 4.1470.60a


3.6270.45b 0.8770.38b 0.9870.39b 5.7770.65b
Ribose-5-P 1.1170.29a 1.2870.33a 0.5570.25a 1.9870.40a
Ribulose-5-P 4.3071.54a 4.3671.52a 4.1871.46a 4.2071.52a

a
Sampled during exponential growth period cultivated in minimal medium.
b
Sampled during exponential growth period cultivated in LBG medium.

Table 5
Specific enzyme activities in cell extracts of B. subtilis, sampled during exponential growth period.

Enzyme activities (nmol min!1 mg!1 protein) 168 RH33 RH33-Prs RH33-PY

PRPP synthetase 4.170.8 3.070.7 10.172.0 10.772.0


Ribose-5-phosphate isomerase B 28.772.5 22.372.5 27.472.5 61.475.5

Table 6
Metabolic characterization of B. subtilis strains during exponential growth period (OD600 0.3–0.6) cultivated in minimal medium.

Parameter 168 RH33 RH33-Prs RH33-PY

!1 !1
Specific glucose uptake rate (mmol g CDW h ) 4.570.4 4.270.4 4.170.4 4.270.4
Specific growth rate (h!1) 0.3170.03 0.2770.02 0.2870.03 0.2770.03
Specific organic acidsa secretion rate (mmol g!1 CDW h!1)
1.6370.31 1.5170.28 1.4770.22 1.3070.22
Specific riboflavin production rate (mmol g!1 CDW h!1) NDb 0.04570.002 0.04770.003 0.05370.003

Results represent the mean values with standard deviations from three independent measurements.
a
Organic acid including acetate, lactate, and pyruvate.
b
Not detected.
ARTICLE IN PRESS
250 S. Shi et al. / Metabolic Engineering 11 (2009) 243–252

As shown in Fig. 1, ribose-5-phosphate can be generated from 10


ribulose-5-phosphate catalyzed by ywlF encoding ribose-5-phos- 9
phate isomerase B. We decided to co-overexpress the prs and 8
ywlF genes simultaneously. In this study, the plasmid containing
7
both prs gene and ywlF gene under the control of the P43

Glucose (g/L)
promoter was introduced into the prs locus of the B. subtilis 6
RH33 chromosome to construct B. subtilis RH33-PY strain. Both 5
PRPP synthetase and ribose-5-phosphate isomerase B enzyme 4
activities were significantly higher than strain RH33, as shown in 3
Table 5, indicating the successful overexpression of prs and ywlF
2
genes.
When both prs and ywlF were overexpressed, the specific cell 1
0
growth rate and glucose uptake rate were almost the same 10
compared with RH33 (Table 6). Meanwhile, RH33-PY had a 9
slightly lower organic acid secretion. The productivity of riboflavin 8
was significantly increased from 0.04570.002 to 0.0537
7
0.003 mmol g!1 CDW h!1. The concentration of intermediate me-

Biomass (0D600)
tabolites PRPP and ribose-5-phosphate were much higher in B. 6
subtilis RH33-PY compared to RH33 strain (Table 4), which 5
implied that the higher yield of riboflavin was attributable to 4
the sufficient supply of PRPP. Although the overexpression of 3
ribose-5-phosphate isomerase B caused more carbon flux to
2
redirect from ribulose-5-phosphate to ribose-5-phosphate, the
concentrations of ribulose-5-phosphate in these strains were 1
almost constant, which indicated the abundant precursor supply 0
provided by central carbon metabolism in B. subtilis RH33 as 250
Riboflavin Concentration (mg/L)

mentioned above. 200


PRPP has dual roles in riboflavin production: one is used as a
precursor, and the other is to bind PurR and depress PurR-
150
regulated genes. To confirm that indeed the increased PRPP pool
resulted in up-regulation of PurR controlled genes in RH33-PY
strain, a quantitative RT-PCR experiment was conducted to 100
measure the transcript abundances of PurR-regulated genes
relative to parental strain RH33. The fold change of the selected 50
PurR-regulated genes, including purE, purD, guaC, purA, glyA, folD,
pbuG, xpt, pbuX, ytiP, and yqhZ, were shown in Fig. 2. As expected, 0
expression of those PurR-regulated genes, including the purine 0 2 4 6 8 10
pathway genes (purE, purD) and the glycine biosynthesis pathway Time (h)
genes (glyA, folD), were induced in RH33-PY (Fig. 2), which was in
Fig. 3. Time-course profiles of biomass, glucose, and riboflavin concentration of B.
accordance with our hypothesis. In B. subtilis RH33-PY, it is
subtilis RH33 (closed triangles), RH33-Prs (closed squares) and RH33-PY (closed
suggested that the derepressed PurR-regulated genes activated circles) in LBG medium during shake flask batch cultivations.
the pathways that supplying precursors for riboflavin
biosynthesis, which in turn led to the increase of riboflavin
Fig. 3 shows the time-course profiles of riboflavin production,
production.
biomass, and glucose concentration exhibited by RH33, RH33-
Prs, and RH33-PY grown in shake flask with LBG medium. The
7 cell growth and glucose uptake profiles of RH33-Prs and RH33-
PY were almost the same as the parental strain RH33. Little effect
on riboflavin production was observed in RH33-Prs compared
Relative gene expression (RH33-RY/RH33)

6
with RH33. However, the RH33-PY mutant exhibited a riboflavin
titer of 3273% higher than RH33 (Fig. 3). Thus, it indicated
5
that the metabolic engineering strategy identified by trans-
criptome analysis successfully improved the trait of riboflavin
4 overproduction in RH33 on a shake flask scale.

3
3.7. Riboflavin production in fed-batch fermentation
2
To investigate the potential of the strain B. subtilis RH33-PY in a
larger scale fermentor, we tested the performance of strain B.
1
subtilis RH33-PY in fed-batch fermentation using the feeding
profile as described in Materials and methods. The cell growth
0
rates were also very similar among the parental strain RH33 and
pur E pur D pur A gly A fol D guaC pbu G xpt pbuX ytiP yqhZ
two engineered strains RH33-Prs and RH33-PY (Fig. 4A).
Genes Compared to parental strain RH33, an average of 25% increase in
Fig. 2. The fold change of selected PurR-regulated transcripts in B. subtilis RH33- riboflavin titer, up to 15 g l!1, was achieved in RH33-PY (Fig. 4B),
PY, relative to RH33. Values represent mean and s.d. (n ¼ 3). while the improvement was only about 6% in strain RH33-Prs
ARTICLE IN PRESS
S. Shi et al. / Metabolic Engineering 11 (2009) 243–252 251

analysis. By co-overexpressing prs and ywlF in RH33, PRPP


140 pool level was significantly increased. This had dual effects:
one was to derepress the purine genes negatively regulated
by PurR (Fig. 2), and the other was to provide more precursors
120
for purine biosynthesis pathways. The strategy developed in
this work allowed for a 25% increase in riboflavin titer, up to
100 15 g l!1, compared to the parental strain during a 5-l fed-batch
Biomass (OD600)

fermentation.
80 An increasing number of publications have demonstrated how
transcriptome analysis has been successfully applied to under-
60 stand and engineering metabolism of the organisms. This trend
will be expected to continue in the post-genomic era. Unfortu-
40 nately, elucidating and application of the mechanism(s) under-
lying the improved performance of the strains developed by
20 random mutagenesis and selection are still limited, partly due to
the complexities of metabolic and regulatory networks encoded in
0 the even simplest bacterial genome. The whole-genome gene
0 6 12 18 24 30 36 42 48 expression analysis, integrated with nucleotide sequencing and
Feed time (h) metabolite measurements, represents a powerful approach to
exploring and exploiting of genomic data, thus greatly facilitating
the identification of novel targets for strain improvement. This
18 systems strategy should be in principle applied for rational
microbial development by deducing the global physiology and
16 regulation of an overproducer strain that is obtained by
Riboflavin concentration (g/L)

14 combinatorial approaches.

12
10 Acknowledgments

8 This research was supported by the National Natural Science


6 Foundation of China (NSFC-20536040), the National Project of Key
Fundamental Research (2007CB707802), the Development Project
4 of Science and Technology of Tianjin (05YFGZGX04500) and
2 Programme of Introducing Talents of Discipline to Universities
(B06006). CapitalBio Co. Ltd., in Beijing was acknowledged for
0 help with the microarray experiment. The microarray data
0 6 12 18 24 30 36 42 48 described in this publication had been deposited in the public
Feed time (h) database (Gene Expression Omnibus, GEO) with accession
number of GSE12873.
Fig. 4. Time-course profiles of biomass (A) and riboflavin concentration (B) of B.
subtilis RH33 (closed triangles), RH33-Prs (closed squares) and RH33-PY (closed
circles) during fed-batch cultivations in a 5-l bioreactor.
Appendix A. Supplementary material
!1
(12.7 g l ). Therefore, in fed-batch fermentation, modulating
PRPP pool by co-overexpressing prs and ywlF successfully Supplementary data associated with this article can be found
improved riboflavin production in B. subtilis while causing little in the online version at doi:10.1016/j.ymben.2009.05.002.
effect on cell growth. These data also suggested that increasing
PRPP pool may be a general strategy to improve riboflavin
production. The alterations of riboflavin production in the three
References
strains have the same trend in the medium tested. Another
successful example for enhancing riboflavin production by
Anagnostopoulos, C., Spizizen, J., 1961. Requirements for transformation in Bacillus
manipulation of the PRPP pool was achieved in Ashbya gossypii subtilis. J. Bacteriol. 81, 741–746.
(Jimenez et al., 2008). Bai, D.M., Zhao, X.M., Li, X.G., Xu, S.M., 2004. Strain improvement of Rhizopus oryzae
In this study, based on known regulatory and metabolic for over-production of L(+)-lactic acid and metabolic flux analysis of mutants.
Biochem. Eng. J. 18, 41–48.
information and new insights generated from transcriptome Bailey, J.E., 1991. Toward a science of metabolic engineering. Science 252,
analysis, combined with nucleotide sequencing and intracellular 1668–1675.
metabolite concentration measurement, metabolic bottleneck and Bresler, S.E., Glazunov, E.A., Chernik, T.P., Schevchenko, T.N., Perumov, D.A., 1973.
Study of riboflavin biosynthesis operon in Bacillus subtilis. Flavinmononucleo-
novel target genes were identified and co-amplified to further tide and flavinadeninedinucleotide as effectors of the riboflavin operon.
enhance production of riboflavin in B. subtilis. Microarray analysis Genetika 9, 84–92.
identified that purine and glycine biosynthesis genes were Choi, J.H., Lee, S.J., Lee, S.J., Lee, S.Y., 2003. Enhanced production of insulin-like
growth factor I fusion protein in Escherichia coli by coexpression of the down-
significantly down-regulated in the riboflavin-producing strain
regulated genes identified by transcriptome profiling. Appl. Environ. Microbiol.
RH-33. These genes were repressed by the global regulator PurR 69, 4737–4742.
whose activity was modulated by small molecular PRPP, which is Crueger, W., Crueger, A., 1984. Biotechnology: A Textbook of Industrial Micro-
also the precursor for purine synthesis. It was inferred that the biology. Sinauer Associates, Sunderland, MA.
Fisher, S.H., Mangasanik, B., 1984. Synthesis of oxaloacetate in Bacillus subtilis
reduced expression of PurR-regulated genes might be caused by a mutants lacking the 2-ketoglutarate dehydrogenase enzymatic complex. J.
low PRPP pool, which was subsequently confirmed by metabolite Bacteriol. 158, 55–62.
ARTICLE IN PRESS
252 S. Shi et al. / Metabolic Engineering 11 (2009) 243–252

Gall, S., Lynch, M.D., Sandoval, N.R., Gill, R.T., 2008. Parallel mapping of genotypes Peberdy, J.F., 1985. Biology of penicillins. In: Demain, A.L., Solomon, N.A. (Eds.),
to phenotypes contributing to overall biological fitness. Metab. Eng. 10, Biology of Industrial Microorganisms. Benjamin-Cummings, Menlo Park, pp.
382–393. 407–431.
Harris, D.M., Westerlaken, I., Schipper, D., van der Krogt, Z.A., Gombert, A.K., Peebles, C.A.M., Hughes, E.H., Shanks, J.V., San, K.Y., 2009. Transcriptional response
Sutherland, J., Raamsdonk, L.M., van den Berg, M.A., Bovenberg, R.A.L., Pronk, of the terpenoid indole alkaloid pathway to the overexpression of ORCA3 along
J.T., Daran, J.M., 2009. Engineering of Penicillium chrysogenum for fermentative with jasmonic acid elicitation of Catharanthus roseus hairy roots over time.
production of a novel carbamoylated cephem antibiotic precursor. Metab. Eng. Metab. Eng. 11, 76–86.
11, 125–137. Perkins, J.B., Pero, J.G., Sloma, A., 1991. Riboflavin overproducing strains of bacteria.
Heefner, D.L., Weaver, C.A., Yarus, M.J., Burdzinski, L.A., Gyure, D.C., Foster, E.W., European Patent 0405370.
1988. Riboflavin producing strains of microorganisms, method for selecting, Perkins, J.B., Sloma, A., Pero, J.G., Hatch, R.T., Hermann, T., Erdenberger, T., 1999.
and method for fermentation. Patent WO 88/09822. Bacterial strains which overproduce riboflavin. US Patent 5925538.
Hoque, M.A., Ushiyama, H., Tomita, M., Shimizu, K., 2005. Dynamic responses of Sakuma, R., Nishina, T., Yamanaka, H., Kamatani, N., Nishioka, K., Maeda, M., Tsuji,
the intracellular metabolite concentrations of the wild type and pykA mutant A., 1991. Phosphoribosylpyrophosphate synthetase in human erythrocytes:
Escherichia coli against pulse addition of glucose or NH3 under those limiting assay and kinetic studies using high-performance liquid chromatography. Clin.
continuous cultures. Biochem. Eng. J. 26, 38–49. Chim. Acta 203, 143–152.
Hove-Jensen, B., Maigaard, M., 1993. Escherichia coli rpiA gene encoding ribose Sambrook, J., Russell, D.W., 2001. Molecular Cloning: A Laboratory Manual, third
phosphate isomerase A. J. Bacteriol. 175, 5628–5635. ed. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY.
Humbelin,
+ M., Griesser, V., Keller, T., Schurter, W., Haiker, M., Hohmann, H.P., Ritz, Sauer, U., Hatzimanikatis, V., Hohmann, H.P., Manneberg, M., Van Loon, A.P.G.M.,
H., Richter, G., Bacher, A., Van Loon, A.P.G.M., 1999. GTP cyclohydrolase II and Bailey, J.E., 1996. Physiology and metabolic fluxes of the wild-type
3,4-dihydroxy-2-butanone 4-phosphate synthase are rate-limiting enzymes in and riboflavin-producing Bacillus subtilis. Appl. Environ. Microbiol. 62,
riboflavin synthesis of an industrial Bacillus subtilis strain used for riboflavin 3687–3696.
production. J. Ind. Microbiol. Biotechnol. 22, 1–7. Saxild, H.H., Brunstedt, K., Nielsen, K.I., Jarmer, H., Nygaard, P., 2001. Definition of
Iida, A., Teshiba, S., Mizobuchi, K., 1993. Identification and characterization of the the Bacillus subtilis PurR Operator Using Genetic and Bioinformatic Tools and
tktB gene encoding a second transketolase in Escherichia coli K-12. J. Bacteriol. Expansion of the PurR Regulon with glyA, guaC, pbuG, xpt-pbuX, yqhZ-folD, and
175, 5375–5383. pbuO. J. Bacteriol. 183, 6175–6183.
Izallalen, M., Mahadevan, R., Burgard, A., Postier, B., Didonato, R., Sun, J., Schilling, Schlupen, C., Santos, M.A., Weber, U., de Graaf, A., Revuelta, J.L., Stahmann, K.P.,
C.H., Lovley, D.R., 2008. Geobacter sulfurreducens strain engineered for 2003. Disruption of the SHM2 gene, encoding one of two serine hydro-
increased rates of respiration. Metab. Eng. 10, 267–275. xymethyltransferase isoenzymes, reduces the flux from glycine to serine in
Jaluria, P., Betenbaugh, M., Konstantopoulos, K., Frank, B., Shiloach, J., 2007. Ashbya gossypii. Biochem. J. 369, 263–273.
Application of microarrays to identify and characterize genes involved in Sindelar, G., Wendisch, V.F., 2007. Improving lysine production by Corynebacterium
attachment dependence in HeLa cells. Metab. Eng. 9, 241–251. glutamicum through DNA microarray-based identification of novel target
Jimenez, A., Santos, M.A., Pompejus, M., Revuelta1, J.L., 2005. Metabolic engineer- genes. Appl. Microbiol. Biotechnol. 76, 677–689.
ing of the purine pathway for riboflavin production in Ashbya gossypii. Appl. Stahmann, K.P., Revuelta, J.L., Seulberger, H., 2000. Three biotechnical processes
Environ. Microbiol. 71, 5743–5751. using Ashbya gossypii, Candida famata, or Bacillus subtilis compete with
Jimenez, A., Santos, M.A., Revuelta, J.L., 2008. Phosphoribosyl pyrophosphate chemical riboflavin production. Appl. Microbiol. Biotechnol. 53, 509–516.
synthetase activity affects growth and riboflavin production in Ashbya gossypii. Stephanopoulos, G., Aristidou, A.A., Nielsen, J., 1998. Metabolic Engineering:
BMC Biotechnol. 8, 67–78. Principles and Methodologies. Academic Press, San Diego.
Kamada, N., Yasuhara, A., Takano, Y., Nakano, T., Ikeda, M., 2001. Effect of Tännler, S., Zamboni, N., Kiraly, C., Aymerich, S., Sauer, U., 2008. Screening of
transketolase modifications on carbon flow to the purine-nucleotide pathway Bacillus subtilis transposon mutants with altered riboflavin production. Metab.
in Corynebacterium ammoniagenes. Appl. Microbiol. Biotechnol. 56, 710–717. Eng. 10, 216–226.
Knorr, B., Schlieker, H., Hohmann, H.P., Weuster-Botz, D., 2007. Scale-down and Vogl, C., Grill, S., Schilling, O., Stulke, J., Mack, M., Stolz, J., 2007. Characterization of
parallel operation of the riboflavin production process with Bacillus subtilis. riboflavin (vitamin B2) transport proteins from Bacillus subtilis and Coryne-
Biochem. Eng. J. 33, 263–274. bacterium glutamicum. J. Bacteriol. 189, 7367–7375.
Kornberg, A., Lieberman, I., Simms, E.S., 1955. Enzymatic synthesis and properties Wang, P.Z., Doi, R.H., 1984. Overlapping promoters transcribed by Bacillus subtilis
of 5-phosphoribosylpyrophosphate. J. Biol. Chem. 215, 389–402. sigma 55 and sigma 37 RNA polymerase holoenzymes during growth and
Kunst, F., Ogasawara, N., Moszer, I., Albertini, A.M., Alloni, G., Azevedo, V., Bertero, stationary phases. J. Biol. Chem. 259, 8619–8625.
M.G., Bessieres, P., Bolotin, A., Borchert, S., et al., 1997. The complete genome Weng, M.L., Nagy, P.L., Zalkin, H., 1995. Identification of the Bacillus subtilis pur
sequence of the Gram-positive bacterium Bacillus subtilis. Nature 390, 249–256. operon repressor. Proc. Natl. Acad. Sci. 92, 7455–7459.
Lee, J.M., Zhang, S., Saha, S., Anna, S.S., Jiang, C., Perkins, J., 2001. RNA expression Wierckx, N.J.P., Ballerstedt, H., de Bont, J.A.M., de Winde, J.H., Ruijssenaars, H.J.,
analysis using an antisense Bacillus subtilis genome array. J. Bacteriol. 183, Wery, J., 2008. Transcriptome analysis of a phenol-producing Pseudomonas
7371–7380. putida S12 construct: genetic and physiological basis for improved production.
Livak, K.J., Schmittgen, T.D., 2001. Analysis of relative gene expression data using real- J. Bacteriol. 190, 2822–2830.
time quantitative PCR and the 2(-Delta Delta C(T)) Method. Methods 25, 402–408. Wu, Q.L., Chen, T., Gan, Y., Chen, X., Zhao, X.M., 2007. Optimization of riboflavin
Mateos, L., Jimenez, A., Revuelta, J.L., Santos, M.A., 2006. Purine biosynthesis, production by recombinant Bacillus subtilis RH44 using statistical designs.
riboflavin production, and trophic-phase span are controlled by a myb-related Appl. Microbiol. Biotechnol. 76, 783–794.
transcription factor in the fungus Ashbya gossypii. Appl. Environ. Microbiol. 72, Zamboni, N., Mouncey, N., Hohmann, H.P., Sauer, U., 2003. Reducing maintenance
5052–5060. metabolism by metabolic engineering of respiration improves riboflavin
Nielsen, J., Olsson, L., 2002. An expanded role for microbial physiology in metabolic production by Bacillus subtilis. Metab. Eng. 5, 49–55.
engineering and functional genomics: moving towards systems biology. FEMS Zamboni, N., Maaheimo, H., Szyperski, T., Hohmann, H.P., Sauer, U., 2004a. The
Yeast Res. 2, 175–181. phosphoenolpyruvate carboxykinase also catalyzes C3 carboxylation at the
Nudler, E., Mironov, A.S., 2004. The riboswitch control of bacterial metabolism. interface of glycolysis and the TCA cycle of Bacillus subtilis. Metab. Eng. 6,
Trends Biochem. Sci. 29, 11–17. 277–284.
Nygaard, P., 1993. Purine and pyrimidine salvage pathways. In: Sonenshein, A.L., Zamboni, N., Fischer, E., Laudert, D., Aymerich, S., Hohmann, H.P., Sauer, U., 2004b.
Hoch, J.A., Losick, R. (Eds.), Bacillus subtilis and Other Gram-Positive Bacteria: The Bacillus subtilis yqjI gene encodes the NADP(+)-dependent 6-P-gluconate
Biochemistry, Physiology, and Molecular Genetics. American Society for dehydrogenase in the pentose phosphate pathway. J. Bacteriol. 186,
Microbiology, Washington, DC, pp. 359–378. 4528–4534.
Park, J.H., Lee, K.H., Kim, T.Y., Lee, S.Y., 2007. Metabolic engineering of Escherichia Zhu, Y.B., Chen, X., Chen, T., Shi, S.B., Zhao, X.M., 2006. Over-expression of glucose
coli for the production of L-valine based on transcriptome analysis and in silico dehydrogenase improves cell growth and riboflavin production in Bacillus
gene knockout simulation. Proc. Natl. Acad. Sci. 104, 7797–7802. subtilis. Biotechnol. Lett. 28, 1667–1672.

Anda mungkin juga menyukai