Download Word

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

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

Document related concepts

Paracrine signalling wikipedia , lookup

Lac operon wikipedia , lookup

Interactome wikipedia , lookup

Lipid signaling wikipedia , lookup

Signal transduction wikipedia , lookup

Gene regulatory network wikipedia , lookup

Fatty acid metabolism wikipedia , lookup

Biochemical cascade wikipedia , lookup

Metalloprotein wikipedia , lookup

Western blot wikipedia , lookup

Endogenous retrovirus wikipedia , lookup

Evolution of metal ions in biological systems wikipedia , lookup

Protein wikipedia , lookup

Point mutation wikipedia , lookup

Genetic code wikipedia , lookup

Enzyme wikipedia , lookup

Protein–protein interaction wikipedia , lookup

Magnesium transporter wikipedia , lookup

Expression vector wikipedia , lookup

Protein structure prediction wikipedia , lookup

Artificial gene synthesis wikipedia , lookup

Biosynthesis wikipedia , lookup

Two-hybrid screening wikipedia , lookup

Metabolism wikipedia , lookup

Proteolysis wikipedia , lookup

Amino acid synthesis wikipedia , lookup

Biochemistry wikipedia , lookup

Transcript
1
Evidence That COG0325 Proteins are involved in PLP Homeostasis
2
Laurence Prunettia, Basma El Yacoubia, Cara R. Schiavona, Ericka Kirkpatricka, Lili Huangb,
3
Marc Baillya*, Mona ElBadawi-Sidhuc, Katherine Harrisona, Jesse F. Gregory 3rd b, Oliver
4
Fiehnc, Andrew D. Hansond and Valérie de Crécy-Lagarda#
5
6
Department of Microbiology and Cell Science, Institute for Food and Agricultural Sciences and
7
Genetic Institute, University of Florida, Gainesville, Florida, USAa; Department of Food Science
8
and Human Nutrition, University of Florida, Gainesville, Florida, USAb; Department of
9
Molecular and Cellular Biology & Genome Center, University of California, Davis, California,
10
USAc; Department of Horticultural Sciences, University of Florida, Gainesville, Florida, USAd
11
12
* Current address: Merck, Palo Alto California, USA
13
L. P. and B. E.-Y. contributed equally to this work
14
#Address correspondence to Valérie de Crécy-Lagard, [email protected]
15
16
Running title: Function of E.coli yggS
17
18
Keywords
19
yggS; COG0325; vitamin B6; PLP protein; PROSC; pyridoxine toxicity; protein of unknown
20
function.
21
22
23
Subject category: Physiology and metabolism
24
Word count
25
Abstract= 195
26
Main text (excluding abstract, and references) =5695
27
28
Abbreviations
29
GSA: Glutamic 5-semialdehyde; P5C: Δ1-pyrroline-5 carboxylic acid;; THF: Tetrahydrofolate;
30
CH2-THF: 5,10-methylene-tetrahydrofolate; Gcv: Glycine-cleavage complex; PLP: Pyridoxal 5’-
31
phosphate; PL: pyridoxal; PN: pyridoxine; PM: pyridoxamine; PNP: Pyridoxine 5’-phosphate;
32
PROSC: PROline Synthase Co-transcribed homolog; ALR: alanine racemase; ODC: ornithine
33
decarboxylase; meth: methionine; dT: thymidine; Cm: chloramphenicol; Amp: ampicillin.; Kan:
34
kanamycin; Tet: tetracycline; aTet: anhydrotetracycline.
35
36
37
ABSTRACT
38
Pyridoxal 5'-phosphate (PLP) is an essential cofactor for nearly 60 Escherichia coli enzymes but
39
is a highly reactive molecule that is toxic in its free form. How PLP levels are regulated and how
40
PLP is delivered to target enzymes are still open questions. The COG0325 protein family
41
belongs to the fold-type III class of PLP enzymes and binds PLP but has no known biochemical
42
activity although it occurs in all kingdoms of life. Various pleiotropic phenotypes of the
43
Escherichia coli COG0325 (yggS) mutant have been reported, some of which were reproduced
44
and extended in this study. Comparative genomic, genetic and metabolic analyses suggest that
45
these phenotypes reflect an imbalance in pyridoxal 5'-phosphate (PLP) homeostasis. The E. coli
46
yggS mutant accumulates the PLP precursor pyridoxine 5'-phosphate (PNP), and is sensitive to
47
an excess of pyridoxine but not of pyridoxal. The pyridoxine toxicity phenotype is
48
complemented by the expression of eukaryotic yggS orthologs. It is also suppressed by the
49
presence of amino acids specifically isoleucine, threonine and leucine suggesting the PLP
50
dependent enzyme transaminase B (IlvE) is affected. These genetic results lay a foundation for
51
future biochemical studies of the role of COG0325 proteins in PLP homeostasis.
52
53
Introduction
54
Pyridoxal 5’-phosphate (PLP) is one of six interconvertible vitamin B6 species (pyridoxal or PL,
55
pyridoxine or PN, pyridoxamine or PM and their 5'-phosphate forms). Enzymes utilizing PLP as
56
cofactor are found in all organisms, catalyze diverse reactions including transamination,
57
decarboxylation, racemization and β- and - elimination, and are mostly associated with amino
58
acid metabolism (Percudani and Peracchi 2009). PLP-enzymes belong to seven structurally
59
distinct families (fold types I-VII) that probably arose independently (Christen and Mehta 2001;
60
Percudani and Peracchi 2003) and encompass 184 enzyme activities. PLP metabolism is
61
particularly relevant to human health because several disorders have been linked to PLP
62
deficiency (Clayton 2006; Halsted 2013; Paul et al. 2013).
63
While the biosynthesis, interconversion and salvage pathways for vitamin B6 species are
64
well characterized (Fitzpatrick et al. 2010; Mooney and Hellmann 2010; Herrero et al. 2011;
65
Mukherjee et al. 2011; Sang et al. 2011; Rueschhoff et al. 2013; Szydlowski et al. 2013), little is
66
known about the regulation of PLP synthesis or about the connection between PLP and general
67
metabolism (Shi et al. 2002; Chen and Xiong 2005; Titiz et al. 2006; Rueschhoff et al. 2013;
68
Vanderschuren et al. 2013). How PLP molecules are delivered to their target enzymes and how
69
the free/bound PLP pool is regulated are also poorly understood (di Salvo et al. 2012). PLP
70
reacts with apo-B6 enzymes by forming an aldimine linkage with the ε-amino group of the active
71
site lysine residue to produce the catalytically active holo-B6 enzyme forms. Alternatively, its
72
highly reactive 4′-aldehyde group can spontaneously form unwanted aldimines with the ε-amino
73
group of lysine residues of non-B6 proteins and with many other amines, and thiazolidine
74
adducts with sulfhydryl groups of molecules such as cysteine, potentially leading to enzyme
75
inactivation (Ohsawa and Gualerzi 1981; Dong and Fromm 1990; Mizushina et al. 2003;
76
Vermeersch et al. 2004) and accumulation of damaged metabolites. For example, PLP reacts
77
through a Knoevenagel condensation with Δ1-pyrroline-5-carboxylate (P5C) (Fig. 1) and Δ1-
78
piperidine-6-carboxylate (P6C) (intermediates in proline and lysine metabolism, respectively)
79
that accumulate in patients deficient in P5C-dehydrogenase and in α-aminoadipic
80
semialdehyde/P6C dehydrogenase, respectively, leading to PLP deficiency (Fig. 1) (Farrant et al.
81
2001; Clayton 2006; Mills et al. 2006).
82
The high chemical reactivity of PLP requires a tight control of the free PLP pool such
83
that an appropriate supply is available to apo-B6 enzymes while undesirable interactions are
84
minimized. It has been proposed that PLP availability is regulated through product feedback
85
inhibition and tight binding to PL kinase (PdxK), PNP oxidase (PdxH) as well as PLP synthase
86
(PdxJ) (Yang and Schirch 2000; Moccand et al. 2011; Ghatge et al. 2012). However, a precise
87
model of PLP homeostasis and how PLP is delivered to target enzymes is still lacking in any
88
organism and is required to understand and manage PLP-related diseases.
89
A candidate for a missing player in PLP homeostasis is the YggS/PROSC/YBL036C
90
family (COG0325). This family belongs to the fold-type III class, along with alanine racemases
91
and certain decarboxylases (Percudani and Peracchi 2009), and has been shown to bind PLP
92
(Eswaramoorthy et al. 2003). Crystal structures of the yeast and Escherichia coli COG0325
93
proteins have been determined [(Eswaramoorthy et al. 2003) and PDB id 1W8G]. Like the N-
94
terminal domain of alanine racemase (ALR) and ornithine decarboxylase (ODC), this protein
95
folds as a TIM barrel with a characteristic long N-terminal helix, and binds PLP in a similar
96
mode. Unlike the dimeric ALR, YBL036C was found to be monomeric, and it was recently
97
reported that the E. coli member of the family, YggS, has no racemase activity towards any of
98
the 20 protein amino acids or their D enantiomers (Ito et al. 2013).
99
Both the broad phylogenetic distribution and the pleiotropic phenotypes linked to the
100
mutations of members of the COG0325 family are suggestive of a core conserved function. In E.
101
coli, the ΔyggS ΔglyA double mutant is not viable on LB medium (Nichols et al. 2011). GlyA
102
(SHMT, serine hydroxymethyltransferase) converts serine to glycine and in the process transfers
103
a hydroxymethyl group to tetrahydrofolate (THF) forming 5,10-methylene-tetrahydrofolate
104
(CH2-THF), the major source of C1 units in the cell (Green et al. 1996) (Fig. 1). GlyA, and the
105
glycine cleavage enzyme system (Gcv), another source of one carbon units, are key to the
106
biosynthesis of purines, thymidine, methionine, and lipids (Fig. 1). Mutants in the THF pathway
107
are sensitive to sulfonamides (targeting FolP) and trimethoprim (targeting FolA) (Nichols et al.
108
2011). Like mutants in the Gcv system, the ΔyggS strain is sensitive to sulfonamides and not to
109
trimethoprim while the serine hydroxymethyltransferase mutant (ΔglyA::KanR) is sensitive to
110
trimethoprim and not to sulfonamides (Nichols et al. 2011). Recently, ΔyggS strains were shown
111
to display altered intracellular amino acid and acetyl coenzyme A (CoA) pools, and to excrete L-
112
valine in the culture medium while accumulating pyruvate, 2-ketobutyrate and 2-aminobutyrate
113
(Ito et al. 2013). Another connection with amino acid metabolism comes from Pseudomonas
114
aeruginosa, whose yggS homolog was proposed to be co-transcribed with proC, the gene
115
encoding P5C reductase that catalyzes the last step of proline synthesis (Fig. 1). This observation
116
gave the PROline Synthase Co-transcribed homolog (PROSC) name to the first studied members
117
of the family including the human homolog, which is expressed in all organs (De Wergifosse et
118
al. 1994). Finally, in yeast, the COG0325 protein YBL036C was induced three-fold in response
119
to the DNA-damaging agent, methyl methanesulphonate (MMS), implying involvement in
120
processes ensuring genetic integrity (Lee et al. 2007).
121
In this work, genetic, biochemical and comparative genomics approaches were combined
122
to show that COG0325 family proteins have a role in PLP homeostasis that could explain the
123
pleiotropic phenotypes of the yggS strain.
124
125
MATERIAL AND METHODS
126
Bioinformatic analyses. The BLAST tools (Altschul et al. 1997) and resources at NCBI
127
(http://www.ncbi.nlm.nih.gov/) were routinely used. Multiple sequence alignments were built
128
using Clustal Omega (Li et al. 2015) or Multalin (Corpet 1988). Protein domain analysis was
129
performed using the Pfam database tools (Finn et al. 2014). Analysis of phylogenetic distribution
130
and physical clustering was performed in the SEED database (Overbeek et al. 2014). Results are
131
available
132
(http://pubseed.theseed.org/SubsysEditor.cgi). The representative genome sets ( set 1 of ~1000
133
genomes and set 2 of ~1500 genomes) were chosen based on phylogenetic diversity
134
previously described (Dailey et al. 2015; Niehaus et al. 2015).The Ortho-MCL database (Chen et
135
al. 2006) was queried for the analysis of COG0325 orthologs in Eukarya and BLASTp (Altschul
136
et al. 1997) searches were performed specifically against sequences of archaeal genomes.
137
Physical clustering was analyzed with the SEED subsystem coloring tool or the SeedViewer
138
Compare Regions tool (Overbeek et al. 2014). The protein association network analysis was
139
performed on the STRING database (string-db.org/) (Szklarczyk et al. 2015). The Enzyme
140
Function Initiative-Enzyme Similarity Tool (EFI-EST) (http://enzymefunction.org/) was used to
141
extract a physical clustering network (Gerlt et al. 2015) as follows. The amino acid sequences of
142
proteins of the IPR011078 InterPro family was extracted to generate a sequence similarity
143
network with an original alignment score threshold of 30 and no restrictions for alignment
in
the
“YggS_2015_Minimal”
subsystem
on
the
public
SEED
server
as
144
lengths. A 65% sequence identity representative node network for IPR011078 was also edited in
145
Cytoscape (Shannon et al. 2003) to produce multiple networks with alignment score thresholds
146
of 40, 50, 60, 70, 80, 90 and 100 by deleting all edges with –log10 (E) values below those
147
thresholds. These seven networks were then used to generate genome neighborhood networks
148
using the EFI Genome Neighborhood Tool with default parameters. The top 11 Pfam protein
149
families in the genome neighborhood networks with alignment scores of 70 and 80 were
150
identified by restricting the network to Pfam protein families with ≥3,000 neighbors. These
151
specific networks were chosen because they are the ones that separate into the most clusters
152
before the network begins to disintegrate into mostly single nodes. The Interactive Tree of Life
153
v2 (ITOL) platform was used to build the gene distribution trees (http://itol.embl.de/index.shtml)
154
(Letunic and Bork 2011). Gene essentiality data was extracted from the Database of Essential
155
Genes (DEG) database (Luo et al. 2014) (http://www.essentialgene.org/). Structures were
156
visualized with the Protein Data Bank (PDB) tools (www.rcsb.org) (Berman et al. 2000). The
157
distribution
158
(http://bioinformatics.unipr.it/B6db) (Percudani and Peracchi 2009). The id-mapping tools of
159
Ecogene
160
(http://www.uniprot.org/) (Li et al. 2015) where routinely used..The positions of regulatory sites
161
for the PdxR family were extracted from RegPrecise 3.0 (http://regprecise.lbl.gov) (Novichkov et
162
al. 2013). The subcellular localization of plant COG0325 proteins was predicted using TargetP
163
(Emanuelsson et al. 2007). Prediction of E. coli promoters was done using RegulonDB (Salgado
164
et al. 2013).
165
Growth conditions and media. Bacteria were grown on Luria Bertani (LB) medium (BD
166
Diagnostics Systems) at 37 °C or on M9 minimal medium (Sambrook et al. 1989) unless
of
PLP
enzymes
in
(http://www.ecogene.org/)
E.
coli
(Zhou
was
and
extracted
Rudd
from
2013)
the
B6
and
database
Uniprot
167
otherwise stated. Growth media were solidified with agar (15 g/l) (BD Diagnostics Systems) for
168
the preparation of plates. Transformation and P1 transductions were performed following
169
standard procedures (Moore 2011). The sensitivity to P1 phage of all recipient strains used was
170
verified. Anhydrotretracycline (aTet, 50 ng/ml), Ampicillin (Amp, 100 µg/ml), Kanamycin (Kan,
171
50 µg/ml), Spectinomycin (Sp, 50 µg/ml) and Chloramphenicol (Cm, 30 µg/ml) were used as
172
appropriate. ilvE avtA::KanR (LSP5001) and ilvE::KanR avtA (LSP5001) were grown on
173
VBE minimal medium 0.5% glucose (w/v) (Whalen and Berg 1982).
174
175
Strain and plasmid constructions. All strains and plasmids used in this study are listed in Table
176
S1 and all oligonucleotides in Table S2. Details of the constructions are described in the
177
supplemental methods.
178
179
Effect of PN on yggS. yggS (VDC6594) cells were grown overnight in 5 ml of M9 glucose
180
0.2% (w/v), diluted 500-fold in 5 ml of M9 medium and plated on M9 glucose 0.2%; 20 l drops
181
of PN at concentrations of 5.9 mM, 590 M, 295 M or 59M were set on the top of the agar
182
and plates were incubated overnight at 37 °C. For complementation experiments, the yggS
183
strain was transformed freshly for every experiment. The plasmids used are listed in Table S1.
184
Cells transformed with pBAD18, pBAD24yggSEc (pBY291.3) were plated on M9 glucose 0.2%
185
(w/v) arabinose 0.2% (w/v), ampicillin 100 g/ml. Cells transformed with pBAD18
186
LOC100191932 were plate on M9 glucose 0.2% (w/v) in presence or in absence of arabinose
187
0.2% (w/v) ampicillin 100 g/ml. Finally, yggS were transformed with pBEY YBL036C
188
(pBEY329.12), pBEY279.1 (empty vector) in the presence or absence of aTet 50 ng/ml.
189
190
Bioscreen growth curves. Cells were grown at 37 °C overnight in M9 supplemented with 0.2%
191
glucose (w/v) and 1 mM glycine, starting from an optical density (OD) of 0.05 at 600 nm. Cells
192
were then diluted 50-fold into sterile 100-well honeycomb plates with cover (Labsystems)
193
containing the corresponding medium: M9 glucose 0.2% (w/v) supplemented or not with glycine
194
1 mM, LB supplemented or not with methionine (0.13 mM) and 0.16 mM thymidine (dT), M9
195
glucose 0.2% (w/v) and glycine 1 mM supplemented or not with 0.4% casamino acids, or with
196
0.13 mM methionine and 0.16 mM thymidine. The OD at 600 nm was measured using a
197
Labsystems Bioscreen C plate reader. Cells were grown at 37 °C with vigorous shaking. Time
198
points were recorded every 30 minutes. For growth in presence of 20 different amino acids, cells
199
were grown in M9 glucose 0.2% (w/v) supplemented with 1 mM glycine and 0.2 mg/ml or 0.1
200
mg/ml of the corresponding amino acids. All growth experiments were performed in 10
201
replicates and standard deviations (SD) were determined
202
203
Extraction and vitamin B6 analysis. B6 species (PLP, PNP, PMP, PL, PN and PM) were
204
extracted from wild-type and yggS pellets (1.0 ml culture of OD600=1.0) in 0.9 ml of 5% (w/v)
205
metaphosphoric acid and 0.1 ml of internal standard 4-deoxypyridoxine (4-dPN, 73 nmol/ml in
206
5% metaphosphoric acid). The suspension was vortexed and sonicated. After centrifugation at
207
10,000 g for 15 min, the supernatant was filtered and a 50-µl sample was taken for HPLC
208
analysis using fluorescence detection (excitation 328 nm, emission 393 nm). The separation was
209
performed on a Microsorb-MV C18 column (150 × 4.6 mm, 5 µm particle size) using a gradient
210
program described by Sampson et al with some modifications (Sampson and O'Connor 1989).
211
Mobile phase A (0.033 M phosphoric acid and 0.008 M 1-octanesulfonic acid, adjusted to pH 2.2
212
with KOH), B (0.033 M phosphoric acid, adjusted to pH 2.2 with KOH) and C (acetonitrile)
213
were used and the gradient program was as follows: 98% A and 2% C for 10 minutes; a linear
214
gradient to 78% B and 22% C for 8 minutes; a linear gradient to 98% A and 2% C for 2 min;
215
column equilibration for 5.0 min with 98% A and 2% C. Total running time was 25 min and the
216
flow rate was 1.2 ml/min. A post-column reagent (1.0 mg/ml sodium bisulfite in 1.0 M
217
potassium phosphate buffer adjusted to pH 7.5 with KOH) was used to enhance fluorescence of
218
PLP. All activities were measured at least three times and standard deviations were determined.
219
220
Chemicals. PNP was provided by the Vanderbilt Institute of Chemical Biology, Chemical
221
Synthesis Core, Vanderbilt University, Nashville, TN 37232‐0412.
222
223
Results
224
yggS is widely distributed and clusters strongly with genes in diverse metabolic and cellular
225
pathways.
226
To gain further insight in the role of the COG0325 family, a subsystem was constructed in the
227
SEED database to capture all COG0325 members (named YggS_2015_Minimal). The yggS gene
228
is widely distributed in Bacteria and Eukaryotes but is quite rare in Archaea where only
229
Methanosarcinales and Aciduliprofundum sp. MAR08-339 harbor yggS orthologs. Members of
230
this family are present almost universally among Bacteria with >90% of the bacterial species in
231
the SEED database containing at least one yggS homolog (see Subsystem and Fig. 2A). A few
232
Bacteria harbor yggS paralogs and these are scattered around the phylogenetic tree (Fig. 2A).
233
COG0325 genes are present in most (but not all) Eukaryotes including yeast, Caenorhabditis
234
elegans, fruit fly, Arabidopsis, Z. mays, zebrafish, chicken, and mammals including humans [see
235
group OG5_127174 in the OrthoMCL database].
236
Physical clustering was first analyzed using the STRING and SEED databases and is
237
summarized in Table 1. The clustering between COG0325 genes and proC observed decades ago
238
(De Wergifosse et al. 1994) was confirmed; clustering between yggS and proC was observed in
239
~20% of the genomes analyzed (3% in the 1000 genome set) (Table 1). However, our data
240
showed that the link between yggS and proline synthesis is not robust (Fig. S1A). First, although
241
proABC genes are quite often clustered, such clusters never include yggS and some Bacteria such
242
as Helicobacter pylori have proAB homologs but lack yggS homologs (Fig. S1A). Second,
243
several bacteria have two proC genes, one that clusters with proAB and another that clusters with
244
yggS (Fig. S1B).
245
Our analysis uncovered other physical clustering associations. The strongest association
246
was with cell division/cell wall related genes (Table 1). The yggS gene is often located near the
247
end of the well-known division and cell wall (dcw) cluster (Tamames et al. 2001) and also
248
clusters independently with two genes also often located at the end of the dcw cluster yggT
249
(YlmG) and sepF (Table 1). Physical clustering associations also linked YggS to PLP salvage
250
and to the PLP dependent enzyme GlyA (Table 1 and Fig. S2). Indeed, yggS genes were found
251
associated with pdxK, glyA and B6 transporter genes in operons predicted to be under the control
252
of various PdxR-type PLP-responsive transcription regulators (Jochmann et al. 2011; Suvurova
253
and Rodionov 2015; Tramonti et al. 2015) (Fig. S1B and Fig. S2). Finally, genes encoding
254
essential ATP-dependent enzymes (with essential lysines) were found to cluster strongly with
255
yggS (Table 1). Examples include the leucyl- and isoleucyl-tRNA synthetase genes that in
256
combination clustered with yggS in ~25-30% of the genomes.
257
258
Vitamin B6 homeostasis defects are observed in the yggS strain.
259
YggS is a PLP-binding protein (Eswaramoorthy et al. 2003), and our comparative genomic
260
analyses linked members of the YggS family to vitamin B6 synthesis (Table 1, Fig. S2). To
261
explore a potential role of this protein family in vitamin B6 homeostasis in vivo, we tested
262
whether the absence of yggS led to phenotypes in conditions of B6 excess. An E.coli K12 yggS
263
derivative VDC6594 was constructed and tested for response to B6 vitamers. Excess PN led to a
264
toxicity ring on minimal medium in the yggS but not in the wild-type strain (Fig. 3, Table 2 and
265
Table 3). PLP, PM or PL were not toxic in any background (data not shown). This PN toxicity
266
ring was very reproducible (Table 2) and was complemented by expressing the E. coli yggS gene
267
in trans (Fig. 3C) and suppressed by the presence of PL (Table 2). The presence of ring, and not
268
a halo, couldsuggest that suppressors appeared. To discriminate between the two hypotheses, we
269
re-isolated yggS cells from the inside or from the outside of the PN toxicity ring and both
270
retained the PN sensitivity phenotype (Fig. S3A). We also showed that overexpressing pdxK led
271
to PN toxicity in the WT strain (Fig. S3B), the absence of yggS did not seem to increase this
272
toxicity (data not shown).
273
The universality of the YggS function first reported by Ito et al. (Ito et al. 2013) was
274
confirmed here, as the PN toxicity phenotype was complemented by expressing the COG0325
275
gene from Z. mays, LOC100191932, in the presence of inducer (Fig. 3D). No complementation
276
was observed in absence of inducer (Fig. 3E). Similar results were found with the COG0325
277
gene from A. thaliana, At1g11930 (Table 2) but in addition it appeared that overexpression of
278
that gene was toxic (data not shown). Finally, it was found that expression of the yeast COG0325
279
gene, YBL036C, also complemented the PN toxicity phenotype (Fig. 3F).
280
As the growth phenotypes were consistent with a role of YggS in vitamin B6
281
homeostasis, we analyzed the B6 pools in the wild-type and yggS strains grown in M9 glucose
282
at the end of the exponential phase (Fig. 4A). The mutant accumulated PNP 71.64 ± 6.10
283
pmol/mg of proteins and no PNP is detected in the wild-type (Fig. 4B). No difference in the PLP
284
levels was observed between the yggS and the wild-type cells. Of note the total pool of PLP is
285
measured here not the free pool. Because the free PLP pool is only a small proportion of the total
286
pool (di Salvo et al. 2011), changes in the free pool will not be detected with this method, hence
287
it still to be determined if the absence of yggS affects the free PLP pool. Methods to measure the
288
PLP-ome or the PLP bound to proteins exists but they are all semi-quantitative (Whittaker et al.
289
2015). We still decided to compare the bound PLP (PLP to proteins) in the WT and yggS
290
strains using an antibody that detects PLP bound to proteins (Whittaker et al. 2015). The soluble
291
fractions of both strains were separated by 2D gels (Fig. S4A and B) and PLP-bound proteins
292
were detected using the anti-PLP antibody. 17 major protein spots were detected in both strains
293
and no major differences were observed (Fig. S4C and S4D). The absence of yggS does not have
294
a major impact on the bound PLP pool, but clearly the method is not quantitative enough to
295
detect small differences.
296
297
YggS is a monomeric PLP protein that has no direct effect on PdxH enzymatic activity.
298
The structure of E. coli YggS shows that PLP is covalently bound to lysine residue 36
299
through a Schiff base linkage (PDB: 1W8G). The role of PLP in YggS quaternary structure or
300
stability has yet to be determined. Recombinant YggS purified by Ni-affinity as described in the
301
supplemental methods and gel filtration chromatography migrated in SDS-PAGE as a single
302
band with a molecular mass of ~30 kDa (Fig. S5A). The monomeric state observed in the crystal
303
structure was confirmed by gel filtration chromatography, which indicated a native molecular
304
mass of 22.6 kDa (data not shown). The presence of PLP in the sample was detected by optical
305
spectroscopy (Fig. S5B); 32% of the purified YggS contained bound PLP.
306
Two forms of YggS were generated: apo-YggS with no detectable bound PLP, and native
307
holo-YggS in which PLP was reincorporated to the purified YggS (82% PLP). The holo- and
308
apo-YggS gave a symmetrical peak upon gel filtration that corresponded to the monomeric state
309
(data not shown). Thus, PLP had no effect on YggS oligomerization. The ability of apo-YggS to
310
bind PLP was confirmed by ITC with a Kd value was 0.37 ±0.49 (SD) M.
311
To determine whether YggS can bind PNP, a comparative analysis of thermal stability of
312
the purified YggS (with 32% of PLP binding sites occupied by PLP) was carried out by
313
thermofluor assay (Ericsson et al. 2006; Lavinder et al. 2009; Phillips and de la Pena 2011) in the
314
absence or presence of PNP or PLP (Fig. S5C and S5D). Only PLP was found to stabilize YggS
315
as the apparent Tm of the YggS was 86 °C with PLP compared to 78 °C without it. In presence of
316
PNP the apparent Tm of YggS was 78 °C. That PNP did not affect the thermal stability of YggS
317
even at high concentration (data not shown) indicates that YggS does not bind PNP.
318
PdxH (pyridoxine 5'-phosphate oxidase) mediates a key step in the PLP biosynthesis
319
pathway using PNP as substrate (Zhao and Winkler 1995; di Salvo et al. 2011) (Fig. 1). Free PLP
320
inhibits both human and E. coli PdxH and interacts with FMN when it is present (Musayev et al.
321
2003). (Fig. 4) A defect of the PNP oxidase activity could lead to the accumulation of PNP in
322
the yggS strain (Fig.4). This hypothesis was tested by measuring PNP oxidase specific activity
323
in the soluble fraction of the wild-type and the yggS strains as described in the supplemental
324
methods. No difference in PdxH specific activity was observed. The specific activities of
325
theyggS and wild-type were 2.9 ±0.02 (SD) and 2.11 ± 0.12 (SD) nmol h-1 mg-1.
326
327
328
329
Pleotropic effects of yggS depletion.
Previous studies have linked the absence of yggS to amino acid pool imbalances,
particularly the secretion of valine in E.coli MG1655 (Ito et al. 2013).
330
A full analysis of the intracellular and extracellular metabolite pools extracted from the
331
wild-type and yggS strains grown in minimal medium was therefore performed and the
332
metabolomes of the two strain compared by gas chromatography time-of-flight mass
333
spectrometry as described in the supplemental methods (Table S3 and Table S4). Among the
334
most striking differences observed were increases in putrescine, N-acetylputrescine, and
335
putrescine-related metabolites such as γ-aminobutyric acid (and its cyclized form γ-
336
butyrolactam) and 5’-deoxy-5’-thiomethyladenosine in the yggS strain (Table S3). Putrescine is
337
the major polyamine in E. coli (Cohen 1998) and the regulation of its catabolism is very
338
complex, being induced by different stresses including nitrogen limitation (Schneider et al.
339
2013); several enzymes in these pathways are PLP-dependent (Samsonova et al. 2003; Kurihara
340
et al. 2005; Samsonova et al. 2005). In the conditions used in our study, we did not observe
341
significant accumulation of valine in the yggS strain or valine release to the medium (Table S4,
342
and Table 3). Nor could valine secretion by the yggS mutant be detected by feeding the valine
343
requirements of the valine auxotrophs LSP5001 or 5002 (data not shown).
344
Another link between YggS, PLP, and amino acid metabolism was observed with the
345
synthetic lethality phenotype of the E. coli ΔyggS::CmR and ΔglyA::KanR alleles on the rich
346
medium LB (Nichols et al. 2011). GlyA is a PLP-dependent enzyme involved in glycine and
347
serine metabolism (Fig. 1). As this phenotype had only been observed in a high throughput
348
experiment, a yggSglyA derivative was constructed in a BW25113 background. The double
349
mutants did not grow on LB (Fig. S6 and Table 3) but grew on M9 glucose 0.2% (w/v)
350
supplemented with 1 mM glycine (Fig. 5A). The ΔglyA strain also showed a growth phenotype
351
in LB (Fig. S6) as previously reported (Nichols et al. 2011) but it was not as severe as the double
352
mutant. It was then shown that the LB sensitivity phenotype is caused by the presence of amino
353
acids in the medium, as addition of casamino acids 0.4% (w/v) (Fig. 5B and Table 3). Casamino
354
acids also affected growth of the glyA strain (VDC6664) but not of the yggS (VDC6594) or
355
wild-type (WT) (BW25113) strains (Fig. 5B). Growth was then followed with all 20 protein L-
356
amino acids (added individually in presence of 1 mM glycine). Out of these 20, threonine (0.84
357
mM) and alanine (2.24 mM) delayed growth of the double mutant compared to the glyA, yggS
358
and wild-type strains (Fig. S7A and S7B) even if the final growth rates were similar (~0.04
359
OD600nm.h-1). The growth defects with single amino acids were not as pronounced as with LB or
360
casamino acids, suggesting additive or synergistic effects. The addition of dT and methionine
361
also partially suppressed the phenotype (Fig. 5C).
362
Finally, because of these links with amino acid metabolism, we tested whether amino
363
acids affected the PN toxicity phenotype of the yggS. We found that the PN toxicity ring
364
observed in the yggS strain (Fig. 5D) was suppressed by the presence of casamino acids (Fig.
365
S8B and Table 3) and that out of the 20 amino acids tested individually, only leucine (Fig. 5D),
366
isoleucine (Fig. 5E) or threonine (Fig. 5F) suppressed the ring of toxicity (Table 3, Fig. S8).
367
Valine could not be tested because it is toxic. This suppression was not due to an increase growth
368
rate as adding L-leucine, L-isoleucine or L-threonine did not affect the growth rate of the yggS
369
mutant (data not shown). Thr is a precursor or Ile and the last step in the synthesis of Ile and Leu
370
is catalyzed by the PLP-dependent transaminase B or IlvE (Fig. 1). The levels of IlvE were
371
analyzed by Western blot and a slight but reproducible increase in levels of IlvE protein in the
372
yggS strain was observed (Fig. S9).
373
The strongest physical clustering observed in our comparative genomic analysis was
374
between yggS genes and cell division and murein synthesis genes (Table 1). Scanning electron
375
microscopy supported this functional link, indeed yggS glyA cells grown in minimal medium
376
were found to be 23% longer than wild-type cells and a few filaments were observed (Fig. S10
377
and Table 3). The morphology and length of yggS or glyA cells were similar to those of wild-type
378
cells and the addition of D-alanine (1 mM), PL (1 mM), or the combination of methionine (0.13
379
mM) and dT (0.16 mM) did not suppress the observed phenotype in the conditions tested (data
380
not shown).
381
Discussion
382
This study clearly connects YggS, the E. coli member of the COG0325 family, with B6
383
metabolism. First, comparative genomic associations firmly linked the COG0325 family to PLP
384
salvage (Fig. S2). Second, we confirmed that YggS binds PLP as previously reported (Ito et al.
385
2013), and determined the Kd value. We also showed that PLP binding does not affect the
386
quaternary structure of the protein. Finally, we show that the E. coli yggS strain accumulates
387
PNP and that high levels of exogenous PN (but not PL, PM or PLP) are toxic to the mutant
388
strain. In the toxicity tests, the formation of a toxicity ring was very reproducible, however the
389
transformedyggS cells could not be stored for more than a week at 4°C without losing the
390
phenotype. Moreover, bacterial growth was observed in the center of the ring of toxicity. We
391
showed that these were not suppressors but further work is needed to understand concentration
392
dependence of the PN sensitivity phenotype.
393
YggS is widely distributed among bacteria, fungi, and eukaryotes and the universality of
394
YggS function is strengthened by the fact that expressing orthologs from yeast and plants
395
complement the PN toxicity phenotype of the E. coli yggS mutant. However, both the cause of
396
PN toxicity and the exact role of YggS in vitamin B6 homeostasis remain to be elucidated. PN
397
neurotoxicity has been previously reported in mammals (rats, dogs and humans) (Schaeppi and
398
Krinke 1982; Albin et al. 1987; Perry et al. 2004) and we show also here that PN is toxic when
399
PdxK is overproduced in E.coli also suggestive of PNP toxicity (Fig. S3B). However, the
400
mechanism of this toxicity is unknown although it was suggested that PNP competitively inhibits
401
PLP-dependent enzymes in the brain (Albin et al. 1987). That PN (or PNP) could inhibit PLP
402
enzymes in E. coli in the absence of YggS fits with the suppression of the PN toxicity phenotype
403
by PL (if this allows to rebalance the PNP/PLP pools) or casamino acids (if this bypasses the
404
need for specific PLP enzymes). One candidate PLP enzyme is IlvE as adding Ile, Val or Thr
405
suppresses the phenotype and as -ketoglutarate a product of IlvE activity was found to be
406
reduced in the metabolomics analysis (Table S3). How YggS would protect from this PN toxicity
407
is not clear. Indeed, YggS does not protect by sequestering PNP as we showed that it does not
408
bind PNP. We found that PdxH specific activity was not affected by the absence of YggS so
409
other type of regulations must occur and further studies are required to elucidate the causes of
410
PNP increase.
411
As summarized Table 3, YggS depletion has a pleotropic effect on metabolism and
412
diverse phenotypes were observed. We propose that most of the phenotypes observed in E. coli
413
in the absence of yggS are caused by lower activities of PLP-dependent enzymes (Fig. 1).
414
Because PLP participates in nearly sixty reactions in E. coli (http://bioinformatics.unipr.it/B6db),
415
physiological studies in strains with defects in PLP metabolism are difficult to interpret, mainly
416
because: 1) several independent phenotypes are observed; 2) these phenotypes can be due to
417
primary and secondary effects; 3) causes cannot be distinguished from effects. Such difficulties
418
are why it took decades to decipher the molecular basis of phenotypes caused by the absence of
419
ridA that, possibly like yggS, perturbs whole cellular networks by inhibiting PLP enzymes
420
(Downs and Ernst 2015). In addition, even if the core function of YggS is conserved, as the
421
cross-kingdom complementations observed here and in previous studies suggest, the network
422
perturbations caused by its absence may vary from one organism to another. For example, YggS
423
is dispensable in most organisms studied so far but is essential in Pseudomonas aeruginosa
424
(Rusmini et al. 2014) and possibly in the pathogenic bacteria Haemophilus influenzae,
425
Helicobacter pylori, Streptococcus pneumoniae and Staphylococcus aureus (data extracted from
426
the DEG database (Luo et al. 2014)). These combinations of issues make the reproducibility and
427
interpretation of physiological and metabolomic data in mutants affected in PLP metabolism
428
very difficult, as discussed below.
429
According to Ito et al., the absence of yggS leads to a 10% decrease of CoA that triggers
430
accumulation of -ketobutyrate and an increase in the L-valine pool (Ito et al. 2013). These
431
results were not reproduced in our study as no significant increase in L-valine was observed and
432
the -ketobutyrate pool was found to decrease and not increase in most conditions. The
433
differences between the two studies could be due to the different genetic backgrounds
434
(BW25133 or MG1655), the different growth temperatures (30 °C or 37 °C) or to the growth
435
phase at which the analyses were made. Indeed, metabolite pools varied greatly with the stage of
436
growth (Table S3). Further analytical studies are required to reconcile these discrepancies.
437
The growth defect of the yggS glyA strain in the presence of exogenous amino acids is a
438
strong phenotype, but remains unexplained. The partial suppression of this phenotype by dT and
439
methionine suggests a link to THF metabolism that fits with the previously reported sensitivity to
440
sulfonamides of a yggS mutant (Nichols et al. 2011). Our data suggested a strong link of YggS to
441
amino acid metabolism, and indeed metabolomics analysis does reveal changes in levels between
442
wild type and yggS mutant for six amino acids. Finally, our metabolomics data suggested that
443
several PLP-dependent enzymes involved in the putrescine metabolism could have been affected
444
in the yggS strain.
445
Another phenotype observed was the cell division defect of the glyA yggS strain. This
446
could be due to lower levels of D-alanine if the PLP-dependent alanine racemase activity is
447
lowered in this background as shown in numerous studies on B6 limitation (Grogan 1988; Kim et
448
al. 2010). As the addition of D-alanine did not suppress the phenotype it could also have more
449
complex causes. Indeed, it was previously reported that cells containing a pdxH mutation placed
450
under PL limitation were elongated and contained nucleoids that could not segregate and this
451
phenotype was not suppressed by PN or D-alanine, suggesting that it was caused by an early cell
452
division defect (Lam and Winkler 1992). The strong clustering of yggS with genes involved in
453
cell division combined with the observed phenotypes point to a regulatory role in this process
454
that will require further investigation.
455
Finally, our comparative genomics data suggest that the strong linkage of yggS with proC
456
is almost certainly due to a function of proC not directly related to proline synthesis. We propose
457
a role of ProC in PLP metabolism, as in humans as the ProC deficiency depletes the PLP pool
458
through the formation of PLP-P5C adducts (Farrant et al. 2001)(Fig. 1). In Bacteria, ProC could
459
be a mechanism to detoxify excess P5C and we are currently exploring this hypothesis.
460
This work has laid the foundation to elucidate the molecular function of the COG0325
461
family. Possibilities include a carrier function to deliver PLP to the target enzymes or a
462
protective function so that PLP does not inactivate essential lysines in proteins. More generally,
463
like the predatory genetic studies of the rid family (Downs and Ernst 2015), our work
464
underscores the power of genetics to uncover, and to start unraveling, complex biochemical
465
processes.
466
467
Acknowledgments
468
We thank S. Shanker and the staff at UF ICBR for DNA sequencing and the ICBR
469
members Cecilia Silva-Sanchez and Sixue Chen for help with 2D-gel analyses. We gratefully
470
acknowledge the staff at UF ICBR for microscopy. We thank the Cambillau laboratory for the E.
471
coli yggS expression plasmid and Robert McKenna, Antonette Bennett, and the University Of
472
Florida Center Of Structural Biology for help with data acquisition and interpretation of the
473
thermal denaturation and titration calorimetry studies. This work was funded by US National
474
Science Foundation (NSF) grants MCB-1153413 and IOS-1025398. MB was a recipient of a
475
postdoctoral fellowship from the Human Frontier Scientific Program.
476
477
478
References
479
480
Albin R. L., Albers J. W., Greenberg H. S., Townsend J. B., Lynn R. B., Burke J. M., Jr.,
481
Alessi A. G. 1987. Acute sensory neuropathy-neuronopathy from pyridoxine overdose.
482
Neurology 37: 1729-1732.
483
Altschul S. F., Madden T. L., Schaffer A. A., Zhang J., Zhang Z., Miller W., Lipman D. J.
484
1997. Gapped BLAST and PSI-BLAST: a new generation of protein database search
485
programs. Nucleic Acids Res 25: 3389-3402.
486
487
Berman H. M., Westbrook J., Feng Z., Gilliland G., Bhat T. N., Weissig H., Shindyalov I.
N., Bourne P. E. 2000. The Protein Data Bank. Nucleic Acids Res 28: 235-242.
488
Chen F., Mackey A. J., Stoeckert C. J., Jr., Roos D. S. 2006. OrthoMCL-DB: querying a
489
comprehensive multi-species collection of ortholog groups. Nucleic Acids Res 34: D363-
490
368.
491
492
493
494
495
496
Chen H., Xiong L. 2005. Pyridoxine is required for post-embryonic root development and
tolerance to osmotic and oxidative stresses. Plant J 44: 396-408.
Christen P., Mehta P. K. 2001. From cofactor to enzymes. The molecular evolution of
pyridoxal-5'-phosphate-dependent enzymes. Chem Rec 1: 436-447.
Clayton P. 2006. B6-responsive disorders: A model of vitamin dependency. J Inherit Metab Dis
29: 317-326.
497
Cohen S. S. 1998. A guide to the polyamines Oxford Univ. Press, New York.
498
Corpet F. 1988. Multiple sequence alignment with hierarchical clustering. Nucleic Aids Res 16:
499
10881-10890.
500
Dailey H. A., Gerdes S., Dailey T. A., Burch J. S., Phillips J. D. 2015. Noncanonical
501
coproporphyrin-dependent bacterial heme biosynthesis pathway that does not use
502
protoporphyrin. Proc Natl Acad Sci U S A 112: 2210-2215.
503
De Wergifosse P., Jacques B., Jonniaux J. L., Purnelle B., Skala J., Goffeau A. 1994. The
504
sequence of a 22.4 kb DNA fragment from the left arm of yeast chromosome II reveals
505
homologues to bacterial proline synthetase and murine alpha-adaptin, as well as a new
506
permease and a DNA-binding protein. Yeast 10: 1489-1496.
507
508
509
510
di Salvo M. L., Contestabile R., Safo M. K. 2011. Vitamin B(6) salvage enzymes: mechanism,
structure and regulation. Biochim Biophys Acta 1814: 1597-1608.
di Salvo M. L., Safo M. K., Contestabile R. 2012. Biomedical aspects of pyridoxal 5'phosphate availability. Front Biosci (Elite Ed) 4: 897-913.
511
Dong Q., Fromm H. J. 1990. Chemical modification of adenylosuccinate synthetase from
512
Escherichia coli by pyridoxal 5'-phosphate. Identification of an active site lysyl residue. J
513
Biol Chem 265: 6235-6240.
514
Downs D. M., Ernst D. C. 2015. From microbiology to cancer biology: the Rid protein family
515
prevents cellular damage caused by endogenously generated reactive nitrogen species.
516
Mol Microbiol 96: 211-219.
517
518
Emanuelsson O., Brunak S., von Heijne G., Nielsen H. 2007. Locating proteins in the cell
using TargetP, SignalP and related tools. Nat Protoc 2: 953-971.
519
Ericsson U. B., Hallberg B. M., Detitta G. T., Dekker N., Nordlund P. 2006. Thermofluor-
520
based high-throughput stability optimization of proteins for structural studies. Anal
521
Biochem 357: 289-298.
522
Eswaramoorthy S., Gerchman S., Graziano V., Kycia H., Studier F. W., Swaminathan S.
523
2003. Structure of a yeast hypothetical protein selected by a structural genomics
524
approach. Acta Crystallogr D Biol Crystallogr 59: 127-135.
525
Farrant R. D., Walker V., Mills G. A., Mellor J. M., Langley G. J. 2001. Pyridoxal phosphate
526
de-activation by pyrroline-5-carboxylic acid:increased risk of vitamin b6 deficiency and
527
seizures in hyperprolinemia type II. J Biol Chem 276: 15107-15116.
528
Finn R. D., Bateman A., Clements J., Coggill P., Eberhardt R. Y., Eddy S. R., Heger A.,
529
Hetherington K., Holm L.& other authors. 2014. Pfam: the protein families database.
530
Nucleic Acids Res 42: D222-230.
531
Fitzpatrick T. B., Amrhein N., Kappes B., Macheroux P., Tews I., Raschle T. 2007. Two
532
independent routes of de novo vitamin B6 biosynthesis: not that different after all.
533
Biochem J 407: 1-13.
534
535
Fitzpatrick T. B., Moccand C., Roux C. 2010. Vitamin B6 biosynthesis: charting the
mechanistic landscape. Chembiochem 11: 1185-1193.
536
Gerlt J. A., Bouvier J. T., Davidson D. B., Imker H. J., Sadkhin B., Slater D. R., Whalen K.
537
L. 2015. Enzyme Function Initiative-Enzyme Similarity Tool (EFI-EST): A web tool for
538
generating protein sequence similarity networks. Biochim Biophys Acta 1854: 1019-1037.
539
Ghatge M. S., Contestabile R., di Salvo M. L., Desai J. V., Gandhi A. K., Camara C. M.,
540
Florio R., González I. N., Parroni A.& other authors. 2012. Pyridoxal-5'-Phosphate
541
Is a slow tight binding inhibitor of E. coli pyridoxal kinase. PLoS ONE 7: e41680.
542
Green J. C., Nichols B. P., Matthews R. G. 1996. Folate biosynthesis, reduction, and
543
polyglutamylation. in Escherichia coli and Salmonella, Cellular and Molecular Biology
544
(ed. FC Neidhart), pp. 665–673. American Society for Microbiology, Washington, DC.
545
546
547
548
Grogan D. W. 1988. Temperature-sensitive murein synthesis in an Escherichia coli pdx mutant
and the role of alanine racemase. Arch Microbiol 150: 363-367.
Halsted C. H. 2013. B-Vitamin dependent methionine metabolism and alcoholic liver disease.
Clin Chem Lab Med 51: 457-465.
549
Herrero S., Gonzalez E., Gillikin J. W., Velez H., Daub M. E. 2011. Identification and
550
characterization of a pyridoxal reductase involved in the vitamin B6 salvage pathway in
551
Arabidopsis. Plant Mol Biol 76: 157-169.
552
Ito T., Iimori J., Takayama S., Moriyama A., Yamauchi A., Hemmi H., Yoshimura T. 2013.
553
A conserved pyridoxal protein that regulates Ile and Val metabolism. J Bacteriol 195:
554
5439-5449.
555
Jochmann N., Gotker S., Tauch A. 2011. Positive transcriptional control of the pyridoxal
556
phosphate biosynthesis genes pdxST by the MocR-type regulator PdxR of
557
Corynebacterium glutamicum ATCC 13032. Microbiology 157: 77-88.
558
Kim J., Kershner J. P., Novikov Y., Shoemaker R. K., Copley S. D. 2010. Three
559
serendipitous pathways in E. coli can bypass a block in pyridoxal-5'-phosphate synthesis.
560
Mol Syst Biol 6: 436.
561
Kurihara S., Oda S., Kato K., Kim H. G., Koyanagi T., Kumagai H., Suzuki H. 2005. A
562
novel putrescine utilization pathway involves gamma-glutamylated intermediates of
563
Escherichia coli K-12. J Biol Chem 280: 4602-4608.
564
Lam H. M., Winkler M. E. 1992. Characterization of the complex pdxH-tyrS operon of
565
Escherichia coli K-12 and pleiotropic phenotypes caused by pdxH insertion mutations. J
566
Bacteriol 174: 6033-6045.
567
Lavinder J. J., Hari S. B., Sullivan B. J., Magliery T. J. 2009. High-throughput thermal
568
scanning: a general, rapid dye-binding thermal shift screen for protein engineering. J Am
569
Chem Soc 131: 3794-3795.
570
Lee M.-W., Kim B.-J., Choi H.-K., Ryu M.-J., Kim S.-B., Kang K.-M., Cho E.-J., Youn H.-
571
D., Huh W.-K.& other authors. 2007. Global protein expression profiling of budding
572
yeast in response to DNA damage. Yeast 24: 145-154.
573
574
Letunic I., Bork P. 2011. Interactive Tree Of Life v2: online annotation and display of
phylogenetic trees made easy. Nucleic Acids Res 39: W475-478.
575
Li W., Cowley A., Uludag M., Gur T., McWilliam H., Squizzato S., Park Y. M., Buso N.,
576
Lopez R. 2015. The EMBL-EBI bioinformatics web and programmatic tools framework.
577
Nucleic Acids Res.
578
Luo H., Lin Y., Gao F., Zhang C. T., Zhang R. 2014. DEG 10, an update of the database of
579
essential genes that includes both protein-coding genes and noncoding genomic elements.
580
Nucleic Acids Res 42: D574-580.
581
Mills P. B., Struys E., Jakobs C., Plecko B., Baxter P., Baumgartner M., Willemsen M. A.
582
A. P., Omran H., Tacke U.& other authors. 2006. Mutations in antiquitin in
583
individuals with pyridoxine-dependent seizures. Nat Med 12: 307-309.
584
Mizushina Y., Xu X., Matsubara K., Murakami C., Kuriyama I., Oshige M., Takemura M.,
585
Kato N., Yoshida H.& other authors. 2003. Pyridoxal 5'-phosphate is a selective
586
inhibitor in vivo of DNA polymerase alpha and epsilon. Biochem Biophys Res Commun
587
312: 1025-1032.
588
589
Moccand C., Kaufmann M., Fitzpatrick T. B. 2011. It takes two to tango: defining an essential
second active site in pyridoxal 5'-phosphate synthase. PLoS ONE 6: e16042.
590
591
592
593
594
595
Mooney S., Hellmann H. 2010. Vitamin B6: Killing two birds with one stone? Phytochemistry
71: 495-501.
Moore S. D. 2011. Assembling new Escherichia coli strains by transduction using phage P1.
Methods Mol Biol 765: 155-169.
Mukherjee T., Hanes J., Tews I., Ealick S. E., Begley T. P. 2011. Pyridoxal phosphate:
biosynthesis and catabolism. Biochim Biophys Acta 1814: 1585-1596.
596
Musayev F. N., Di Salvo M. L., Ko T. P., Schirch V., Safo M. K. 2003. Structure and
597
properties of recombinant human pyridoxine 5'-phosphate oxidase. Protein Sci 12: 1455-
598
1463.
599
Nichols R. J., Sen S., Choo Y. J., Beltrao P., Zietek M., Chaba R., Lee S., Kazmierczak K.
600
M., Lee K. J.& other authors. 2011. Phenotypic landscape of a bacterial cell. Cell 144:
601
143-156.
602
Niehaus T. D., Gerdes S., Hodge-Hanson K., Zhukov A., Cooper A. J., ElBadawi-Sidhu M.,
603
Fiehn O., Downs D. M., Hanson A. D. 2015. Genomic and experimental evidence for
604
multiple metabolic functions in the RidA/YjgF/YER057c/UK114 (Rid) protein family.
605
BMC Genomics 16: 382.
606
Novichkov P. S., Kazakov A. E., Ravcheev D. A., Leyn S. A., Kovaleva G. Y., Sutormin R.
607
A., Kazanov M. D., Riehl W., Arkin A. P.& other authors. 2013. RegPrecise 3.0--a
608
resource for genome-scale exploration of transcriptional regulation in bacteria. BMC
609
Genomics 14: 745.
610
Ohsawa H., Gualerzi C. 1981. Structure-function relationship in Escherichia coli initiation
611
factors. Identification of a lysine residue in the ribosomal binding site of initiation factor
612
by site-specific chemical modification with pyridoxal phosphate. J Biol Chem 256: 4905-
613
4912.
614
Overbeek R., Olson R., Pusch G. D., Olsen G. J., Davis J. J., Disz T., Edwards R. A.,
615
Gerdes S., Parrello B.& other authors. 2014. The SEED and the Rapid Annotation of
616
microbial genomes using Subsystems Technology (RAST). Nucleic Acids Res 42: D206-
617
214.
618
619
620
621
Paul L., Ueland P. M., Selhub J. 2013. Mechanistic perspective on the relationship between
pyridoxal 5'-phosphate and inflammation. Nutr Rev 71: 239-244.
Percudani R., Peracchi A. 2003. A genomic overview of pyridoxal-phosphate-dependent
enzymes. EMBO Rep 4: 850-854.
622
Percudani R., Peracchi A. 2009. The B6 database: a tool for the description and classification
623
of vitamin B6-dependent enzymatic activities and of the corresponding protein families.
624
BMC Bioinformatics 10: 273.
625
Perry T. A., Weerasuriya A., Mouton P. R., Holloway H. W., Greig N. H. 2004. Pyridoxine-
626
induced toxicity in rats: a stereological quantification of the sensory neuropathy. Exp
627
Neurol 190: 133-144.
628
Phillips K., de la Pena A. H. 2011. The combined use of the Thermofluor assay and ThermoQ
629
analytical software for the determination of protein stability and buffer optimization as an
630
aid in protein crystallization. Curr Protoc Mol Biol Chapter 10: Unit10 28.
631
Rueschhoff E. E., Gillikin J. W., Sederoff H. W., Daub M. E. 2013. The SOS4 pyridoxal
632
kinase is required for maintenance of vitamin B6-mediated processes in chloroplasts.
633
Plant Physiol Biochem 63: 281-291.
634
Rusmini R., Vecchietti D., Macchi R., Vidal-Aroca F., Bertoni G. 2014. A shotgun antisense
635
approach to the identification of novel essential genes in Pseudomonas aeruginosa. BMC
636
Microbiol 14: 24.
637
Salgado H., Peralta-Gil M., Gama-Castro S., Santos-Zavaleta A., Muniz-Rascado L.,
638
Garcia-Sotelo J. S., Weiss V., Solano-Lira H., Martinez-Flores I.& other authors.
639
2013. RegulonDB v8.0: omics data sets, evolutionary conservation, regulatory phrases,
640
cross-validated gold standards and more. Nucleic Acids Res 41: D203-213.
641
642
643
644
645
646
Sambrook J. E., Fritsch F., Maniatis T. 1989. Molecular cloning : A laboratory manual, 2nd
edn. Cold Spring Harbor Laboratory Press, Cold Spring Harbor.
Sampson D. A., O'Connor D. K. 1989. Response of B-6 vitamers in plasma, erythrocytes and
tissues to vitamin B-6 depletion and repletion in the rat. J Nutr 119: 1940-1948.
Samsonova N. N., Smirnov S. V., Altman I. B., Ptitsyn L. R. 2003. Molecular cloning and
characterization of Escherichia coli K12 ygjG gene. BMC Microbiol 3: 2.
647
Samsonova N. N., Smirnov S. V., Novikova A. E., Ptitsyn L. R. 2005. Identification of
648
Escherichia coli K12 YdcW protein as a gamma-aminobutyraldehyde dehydrogenase.
649
FEBS Lett 579: 4107-4112.
650
Sang Y., Locy R. D., Goertzen L. R., Rashotte A. M., Si Y., Kang K., Singh N. K. 2011.
651
Expression, in vivo localization and phylogenetic analysis of a pyridoxine 5'-phosphate
652
oxidase in Arabidopsis thaliana. Plant Physiol Biochem 49: 88-95.
653
Schaeppi U., Krinke G. 1982. Pyridoxine neuropathy: correlation of functional tests and
654
neuropathology in beagle dogs treated with large doses of vitamin B6. Agents Actions 12:
655
575-582.
656
657
Schneider B. L., Hernandez V. J., Reitzer L. 2013. Putrescine catabolism is a metabolic
response to several stresses in Escherichia coli. Mol Microbiol 88: 537-550.
658
Shannon P., Markiel A., Ozier O., Baliga N. S., Wang J. T., Ramage D., Amin N.,
659
Schwikowski B., Ideker T. 2003. Cytoscape: a software environment for integrated
660
models of biomolecular interaction networks. Genome Res 13: 2498-2504.
661
Shi H., Xiong L., Stevenson B., Lu T., Zhu J. K. 2002. The Arabidopsis salt overly sensitive 4
662
mutants uncover a critical role for vitamin B6 in plant salt tolerance. Plant Cell 14: 575-
663
588.
664
Suvurova I. A., Rodionov D. A. 2015. MocR ASK RODIONOV. submitted.
665
Szklarczyk D., Franceschini A., Wyder S., Forslund K., Heller D., Huerta-Cepas J.,
666
Simonovic M., Roth A., Santos A.& other authors. 2015. STRING v10: protein-
667
protein interaction networks, integrated over the tree of life. Nucleic Acids Res 43: D447-
668
452.
669
670
671
672
Szydlowski N., Burkle L., Pourcel L., Moulin M., Stolz J., Fitzpatrick T. B. 2013. Recycling
of pyridoxine (vitamin B6) by PUP1 in Arabidopsis. Plant J 75: 40-52.
Tamames J., Gonzalez-Moreno M., Mingorance J., Valencia A., Vicente M. 2001. Bringing
gene order into bacterial shape. Trends Genet 17: 124-126.
673
Titiz O., Tambasco-Studart M., Warzych E., Apel K., Amrhein N., Laloi C., Fitzpatrick T.
674
B. 2006. PDX1 is essential for vitamin B6 biosynthesis, development and stress tolerance
675
in Arabidopsis. Plant J 48: 933-946.
676
Tramonti A., Fiascarelli A., Milano T., di Salvo M. L., Nogues I., Pascarella S.,
677
Contestabile R. 2015. Molecular mechanism of PdxR - a transcriptional activator
678
involved in the regulation of vitamin B biosynthesis in the probiotic bacterium Bacillus
679
clausii. FEBS J 282: 2966-2984.
680
Vanderschuren H., Boycheva S., Li K. T., Szydlowski N., Gruissem W., Fitzpatrick T. B.
681
2013. Strategies for vitamin B6 biofortification of plants: a dual role as a micronutrient
682
and a stress protectant. Front Plant Sci 4: 143.
683
Vermeersch J. J., Christmann-Franck S., Karabashyan L. V., Fermandjian S., Mirambeau
684
G., Der Garabedian P. A. 2004. Pyridoxal 5'-phosphate inactivates DNA topoisomerase
685
IB by modifying the lysine general acid. Nucleic Acids Res 32: 5649-5657.
686
687
688
689
690
691
Whalen W. A., Berg C. M. 1982. Analysis of an avtA::Mu d1(Ap lac) mutant: metabolic role of
transaminase C. J Bacteriol 150: 739-746.
Whittaker M. M., Penmatsa A., Whittaker J. W. 2015. The Mtm1p carrier and pyridoxal 5'phosphate cofactor trafficking in yeast mitochondria. Arch Biochem Biophys 568: 64-70.
Yang E. S., Schirch V. 2000. Tight binding of pyridoxal 5'-phosphate to recombinant
Escherichia coli pyridoxine 5'-phosphate oxidase. Arch Biochem Biophys 377: 109-114.
692
Zhao G., Winkler M. E. 1995. Kinetic limitation and cellular amount of pyridoxine
693
(pyridoxamine) 5'-phosphate oxidase of Escherichia coli K-12. J Bacteriol 177: 883-891.
694
695
696
697
Zhou J., Rudd K. E. 2013. EcoGene 3.0. Nucleic Acids Res 41: D613-624.
698
Table 1. Summary of physical clustering analysis. Set 1: 1536 representative genome set. Set
699
2: 1000 representative genome set. Total: all 11411 organisms in PubSeed at the time of the
700
analysis. Genes encoding for: * PLP enzymes; proteins with
701
bind GTP or ATP. #no current name for this gene.
&
Essential lysines; proteins that
$
702
Gene
%Set1 %Set2 %Total
Function
Pyrroline-5-carboxylate reductase (EC
1.5.1.2)
SepF, FtsZ-interacting protein related
to cell division
Cell division initiation protein
DivIVA
COG0762, cell division protein
YlmG/Ycf19 (putative), YggT family
Septum formation protein Maf
Cell division protein FtsZ (EC
3.4.24.-)
SepF, FtsZ-interacting protein related
to cell division
UDP-N-acetylglucosamine--Nacetylmuramyl-(pentapeptide)
pyrophosphoryl-undecaprenol Nacetylglucosamine transferase (EC
2.4.1.227)
proC
21.7
3.0
18.4
sepF
22.8
0.8
24.3
divIVA
20.1
1.8
26.1
yggT
19.2
0.5
36.3
maf
2.0
1.5
0.9
ftsZ$
17.3
0.1
29.0
sepF
22.8
0.8
24.3
murG
11.6
0.1
18.0
pilT$
19.2
0.8
30.9
lsp
14.8
1.4
13.5
pdxR
1.2
0.4
0.8
pdxK$
0.9
PLP
0.7
transporter#
0.3
0.5
Lipoprotein signal peptidase (EC
3.4.23.36)
Predicted transcriptional regulator of
pyridoxine metabolism
Pyridoxal kinase (EC 2.7.1.35)
0.4
0.5
B6 transport protein
yqgF
12.0
0.2
23.5
ileS&$
16.7
0.3
25.6
leuS&$
7.6
6.4
5.5
rluD
7.3
0.1
8.7
Twitching motility protein PilT
YqgF, involved in pre-16S RNA
processing
Isoleucyl-tRNA synthetase (EC
6.1.1.5)
Leucyl-tRNA synthetase (EC 6.1.1.4)
Ribosomal large subunit
pseudouridine synthase D (EC
4.2.1.70)
Process
Proline
synthesis
Cell division
Cell division
Cell division
Cell division
Cell division
Cell division
Cell wall
synthesis
Surface
motility
Secretion
PLP salvage
PLP salvage
PLP salvage
Translation
Translation
Translation
Translation
703
704
rdgB
12.0
ND
23.1
yggW
12.4
0.6
23.3
yqgE
11.1
0.3
22.8
topoI
1.3
0.9
0.4
glyA*
0.8
0.1
0.5
aOT3*
0.3
0.5
0.3
Nucleoside 5-triphosphatase RdgB
(dHAPTP, dITP, XTP-specific) (EC
3.6.1.15)
Radical SAM family enzyme, similar
to coproporphyrinogen III oxidase,
oxygen-independent, clustered with
nucleoside-triphosphatase RdgB
UPF0301 protein YqgE
DNA topoisomerase I (EC 5.99.1.2)
Serine hydroxymethyltransferase (EC
2.1.2.1)
Acetylornithine aminotransferase (EC
2.6.1.11)
Nucleotide
metabolism
?
?
Replication
Amino acid
metabolism
Amino acid
metabolism
705
Table 2. Pyridoxine toxicity phenotypes on M9 glucose 0.2% (w/v). % Antibiotics Cm or Amp
706
were added when required; the three plasmids were used as controls and gave similar results.
707
*Arabinose 0.2 % was added. & plate was supplemented with aTet 50 ng/ml.
Strain
Wild-type BW25113
glyA (VDC6664)
yggS pBAD18/pBAD24/pBAD33%*
yggS pBAD24 yggSEC*
yggS pBAD18: LOC100191932*
yggS pBAD18: At1g11930
yggS pBEY YBL036C &
yggS pBAD18 + PL 50 M*
708
709
710
Diameter of the zone of inhibition
(cm)
0.1 mg/ml
1 mg/ml PN
PN
None
None
None
None
4.15 ±0.05
5.25±0.1
None
None
None
None
None
None
None
None
None
None
711
712
Table 3. Phenotypes discussed in this study. M9 Casa is M9 glucose 0.2% (w/v) supplemented
713
with 0.4% (w/v) casamino acids. M9 is M9 glucose 0.2% (w/v) supplemented with isoleucine 50
714
g/ml, leucine 30g/ml , or threonine 30g/ml. $metabolomics data presented in Table 1 and
715
Table S3. *in the conditions used in this study, no significant accumulation of valine was
716
observed in the yggS strain. nd is non determined; - : no phenotype observed; + : phenotype
717
observed. Leu: L-leucine; Ile: L-isoleucine; Thr: L-threonine.
Phenotype
PN sensitivity
PN sensitivity on M9 Casa, M9 Leu, Ile or
Thr
Growth on LB
Growth on M9 Casa
Valine secretion
Longer cells
718
719
720
glyA
-
yggS
+
yggS glyA
+
nd
-
nd
+
+
nd
-
+
+
-$*
-
nd
+
721
Figures legends
722
Figure 1. PLP synthesis and a subset of pathways influenced by PLP levels in E.coli. In E.
723
coli PLP can be synthesized via the de novo PdxFBAJ dependent pathway or salvaged using
724
PdxK, PdxY and PdxH (see review in (Fitzpatrick et al. 2007)). PLP is a cofactor for enzymes
725
involved in cell wall synthesis, amino acids, and one-carbon metabolism. P5C can sequester free
726
PLP.
727
Tetrahydrofolate; CH2-THF: 5,10-methylene-tetrahydrofolate; Gcv: Glycine-cleavage complex;
728
TA: L-threonine aldolase. P5C is formed spontaneously and reversibly from GSA. Dashed
729
arrows indicate that several enzymes are required for a conversion. Proteins in bold with an
730
asterisk are enzymes that require PLP. This figure is based on E. coli literature (see introduction)
731
except for formation of the P5C-PLP adduct that has been reported in human but not E. coli to
732
date Abbreviations for enzymes names are given in the text.
733
Figure 1. PLP synthesis and a subset of pathways influenced by PLP levels in E.coli. In E.
734
coli PLP can be synthesized via the de novo PdxFBAJ dependent pathway or salvaged using
735
PdxK, and PdxH (see review in (Fitzpatrick et al. 2007)). PLP is a cofactor for enzymes involved
736
in cell wall synthesis, amino acids, and one-carbon metabolism. P5C can sequester free PLP.
737
GSA: Glutamic 5-semialdehyde; P5C: Δ1-pyrroline-5 carboxylic acid; THF: Tetrahydrofolate;
738
CH2-THF: 5,10-methylene-tetrahydrofolate; Gcv: Glycine-cleavage complex; TA: L-threonine
739
aldolase; TyrB: aromatic amino acid transferase. P5C is formed spontaneously and reversibly
740
from GSA. Dashed arrows indicate that several enzymes are required for a conversion. Proteins
741
in bold with an asterisk are enzymes that require PLP. This figure is based on E. coli literature
742
(see introduction) except for formation of the P5C-PLP adduct that has been reported in human
743
but not E. coli to date Abbreviations for enzymes names are given in the text.
GSA:
Glutamic
5-semialdehyde;
P5C:
Δ1-pyrroline-5
carboxylic
acid;
THF:
744
Figure 2. Distribution of yggS genes among bacteria. (A) The tree was constructed in iTol
745
(itol.embl.de/). Tree branches are colored by phylum. The presence of yggS homologs are
746
displayed in rings around the tree. The presence of at least one yggS homolog in the specific
747
organism in the tree is shown by the innermost (blue) ring. The presence of two or three yggS
748
paralogs in one genome are shown by the middle (green) ring and the outer (orange) ring
749
respectively. (B) Physical clustering of yggS genes in Pseudomonas sp. GM79 and
750
Rhodopseudomonas palustris TIE-1 with PLP synthesis, salvage genes and PLP dependent
751
enzymes. The red arrows indicate predicted PdxR binding sites.
752
753
Figure 3. Toxic effect of PN on the E. coli yggS strain and the universality of YggS
754
function. The cells were plated as described in the material and methods section. (A) yggS
755
(VDC6594) lawn and drops of PN at concentrations of 5.9 mM (a), 590 M (b), 295 M (c) or
756
59M (d); PN at the concentrations (a) or (b) were used in experiments (B-F). (B) yggS
757
pBAD18 lawn; in presence of arabinose 0.2% (C) yggSin pBAD24yggSEc (pBY291.3) lawn in
758
presence of arabinose 0.2%; (D)yggS transformed with pBAD18 LOC100191932 (pBY298.7)
759
lawn in presence of arabinose 0.2%; (E) yggS transformed pBAD18 LOC100191932 in absence
760
of arabinose 0.2%; (F) yggS transformed with pBEY YBL036C (pBEY329.12) in presence of
761
aTet 50 ng/ml. The arrows indicate the presence of the ring of toxicity.
762
763
Figure 4. Analysis of B6 derivatives in wild-type and yggS E. coli. The B6 vitamers were
764
analyzed by HPLC with fluorometric detection. (A) B6 vitamer profile; 4-dPN is the internal
765
standard 4-deoxypyridoxine. (B) Determination of PNP, PMP, PLP, concentrations (pmol/mg
766
protein) in wild-type (dark gray) and yggS(light gray) strains.
767
768
Figure 5. Effects of amino acids in a yggS background. Wild-type (yellow), yggS
769
(VDC6594) (blue), glyA (orange), and yggS glyA (grey) strains were grown in the Bioscreen
770
C at 37 °C in different media. (M9 glucose 0.2% (w/v) supplemented with 1 mM glycine;
771
(M9 glucose 0.2% (w/v) supplemented with 1 mM glycine and 0.4% casamino acids; (C) LB
772
supplemented with methionine (0.13 mM) and 0.16 mM dT. The average of 10 independent
773
growth curves was plotted, and errors bars represent standard deviations. PN at the
774
concentrations 5.9 mM (a), 590 M (b) were used in experiments (D-F). (D) yggSin pBAD33
775
lawn on M9 supplemented with L-leucine 30 g/mL. (E) yggSin pBAD33 lawn on M9
776
supplemented with L-isoleucine 50 g/mL; (F) (E) yggSin pBAD33 lawn on M9 supplemented
777
with L-threonine 30 g/mL. Rings of toxicity were observed for yggSin pBAD33 lawn on M9
778
(Fig. S7A) for PN at concentrations of 5.9 mM (a), 590 M (b). M9 was supplemented with each
779
20 amino acids individually. Any other amino acid than L-leucine, L-isoleucine and L-threonine
780
gave the same phenotype as Fig. S7C. dT: thymidine; Met: methionine.
781
.
782
783
784