Accepted Article

1 2

Mannitol-1-phosphate dehydrogenases/phosphatases:

A family of novel bifunctional enzymes for bacterial adaptation to osmotic stress

1

3 4

Miriam Sand1, Marta Rodrigues2, José M. González3, Valérie de Crécy-Lagard4,

5

Helena Santos2, Volker Müller1 and Beate Averhoff1*

6 7

1

8

Johann Wolfgang Goethe University Frankfurt am Main, Max-von-Laue-Str. 9, 60438

9

Frankfurt, Germany

Molecular Microbiology & Bioenergetics, Institute of Molecular Biosciences,

10

2

11

Universidade Nova de Lisboa, Av. da República-EAN, 2780-157 Oeiras, Portugal

12

3

13

Spain

14

4

15

32611, USA

Cell Physiology and NMR Lab, Instituto de Tecnologia Química e Biológica,

Department of Microbiology, University of La Laguna, ES-38206 La Laguna, Tenerife,

Department of Microbiology and Cell Science, University of Florida, Gainesville, FL

16 17

Running title: A bifunctional Mtl-1-P dehydrogenases/phosphatase

18 19

*Corresponding author: Beate Averhoff, Molecular Microbiology & Bioenergetics,

20

Institute of Molecular Biosciences, Johann Wolfgang Goethe University Frankfurt/Main,

21

Max-von-Laue-Str. 9, 60438 Frankfurt, Germany. Phone: +49-69-79829509, Fax: +49-69-

22

79829306. E-mail: [email protected]

23 24

This article has been accepted for publication and undergone full peer review but has not been through the copyediting, typesetting, pagination and proofreading process, which may lead to differences between this version and the Version of Record. Please cite this article as doi: 10.1111/1462-2920.12503 This article is protected by copyright. All rights reserved.

2

Accepted Article

1 2

Summary

3 4

The nutritionally versatile soil bacterium Acinetobacter baylyi ADP1 copes with salt

5

stress by the accumulation of compatible solutes, a strategy that is widespread in

6

nature. This bacterium synthesizes the sugar alcohol mannitol de novo in response to

7

osmotic stress. In a previous study, we identified MtlD, a mannitol-1-phosphate

8

dehydrogenase, which is essential for mannitol biosynthesis and which catalyzes the

9

first step in mannitol biosynthesis, the reduction of fructose-6-phosphate (F-6-P) to

10

the intermediate mannitol-1-phosphate (Mtl-1-P). Until now, the identity of the

11

second enzyme, the phosphatase that catalyzes the dephosphorylation of Mtl-1-P to

12

mannitol, was elusive. Here we show that MtlD has a unique sequence among known

13

mannitol-1-phosphate dehydrogenases with a haloacid dehalogenases (HAD)-like

14

phosphatase domain at the N-terminus. This domain is indeed shown to have a

15

phosphatase activity. Phosphatase activity is strictly Mg2+ dependent. NMR analysis

16

revealed that purified MtlD catalyzes not only reduction of F-6-P but also

17

dephosphorylation of Mtl-1-P. MtlD of A. baylyi is the first bifunctional enzyme of

18

mannitol biosynthesis that combines Mtl-1-P dehydrogenase and phosphatase

19

activities in a single polypeptide chain. Bioinformatic analysis revealed that the

20

bifunctional enzyme is wide-spread among Acinetobacter strains but only rarely

21

present in other phylogenetic tribes.

22 23 24 25

This article is protected by copyright. All rights reserved.

3

Accepted Article

1 2

Introduction

3 4

Water is the most important compound on earth and its availability is the prerequisite of

5

life. As cellular membranes are permeable to water, an increase of the external osmolality

6

or desiccation will lead to a loss of water, followed by shrinkage and finally cell death if no

7

countermeasures are taken (Wood et al., 2001). Living cells respond to a decrease in water

8

activities in two ways: halophilic aerobic archaea (halobacteria) accumulate potassium

9

chloride up to molar concentrations in the cells. This requires an adapted cellular

10

machinery that has to work at molar KCl concentrations and restricts growth to high salt

11

concentrations (Galinski and Trüper, 1994; Roeßler and Müller, 2001). More flexible and

12

thus widespread in archaea, bacteria and eukaryotes is the accumulation of small organic

13

molecules that are compatible with the cellular metabolism and thus, they are termed

14

„compatible solutes“, which can be divided in different classes such as polyols, sugars,

15

amino acids and derivatives thereof (da Costa et al., 1998).

16

The metabolically versatile and widespread aerobic soil bacterium Acinetobacter

17

baylyi responds to increasing salinities with the uptake of glycine betaine or its precursor

18

choline from the medium (Sand et al., 2011; Sand et al., 2013b). If glycine betaine is not

19

available it synthesizes glutamate and mannitol de novo (Sand et al., 2013a). The finding

20

of the polyol mannitol as compatible solute was rather surprising, since it had hitherto only

21

been found in eukaryotes such as plants, algae and fungi, and in one bacterial species,

22

Pseudomonas putida, where it is produced together with N-acetylglutaminylglutamine

23

amide (Kets et al., 1996).

24

Mannitol is the most abundant sugar alcohol in nature. It serves as carbon and

25

energy source or radical scavenger and in plants and fungi it is well known as compatible

This article is protected by copyright. All rights reserved.

4 solute (Jennings, 1984; Stoop et al., 1996). In lactic acid bacteria reduction of fructose to

2

mannitol is used to reoxidize NADH produced during glycolysis (Wisselink et al., 2002).

3

However, some homofermentative lactic acid bacteria synthesize mannitol from the

4

glycolytic intermediate, F-6-P, to yield Mtl-1-P that is subsequently dephosphorylated to

5

mannitol (Neves et al., 2000). This pathway has been used to engineer mannitol producing

6

strains by altering the pattern of endproducts (Gaspar et al., 2011). Biotechnological

7

production of mannitol is of high interest since mannitol is widely used in the food,

8

pharmaceutical, medical and chemical industries.

Accepted Article

1

9

So far there are three known routes for the biosynthesis of mannitol in nature. One

10

is the direct reduction of fructose to mannitol by a mannitol-2-dehydrogenase. This

11

enzyme is found in heterofermentative lactic acid bacteria (Wisselink et al., 2002). In a

12

second route mannose-6-phosphate is converted by a mannose-6-phosphate reductase and a

13

phosphatase to mannitol (Rumpho et al., 1983; Loescher et al., 1992). The third

14

mechanism involves the reduction of F-6-P to mannitol via the intermediate Mtl-1-P. This

15

pathway requires two distinct enzymes, a Mtl-1-P dehydrogenase catalyzing the first step

16

and a Mtl-1-P phosphatase catalyzing the second step. This mannitol production pathway

17

is found in homofermentative lactic acid bacteria or algae for instances (Neves et al., 2000;

18

Wisselink et al., 2002; Iwamoto et al., 2003).

19

The biosynthetic route for the production of mannitol as compatible solute in A.

20

baylyi or P. putida was unknown until recently. NMR analysis using A. baylyi revealed

21

Mtl-1-P as intermediate. A rather unusual NADPH-dependent, salt-induced and salt-

22

dependent Mtl-1-P dehydrogenase MtlD (ACIAD1672) catalyzing the reaction F-6-P +

23

NADPH  Mtl-1-P + NADP+ was discovered. Mtl-1-P was further dephosphorylated

24

leading to mannitol (Sand et al., 2013a), but the phosphatase catalyzing the second reaction

25

in this two-step process of mannitol biosynthesis remained to be identified. We report here

This article is protected by copyright. All rights reserved.

5 that MtlD of A. baylyi is a unique bifunctional enzyme mediating both, dehydrogenase and

2

phosphatase activity thus converting F-6-P to mannitol. These data provide first insights

3

into a novel class of Mtl-1-P dehydrogenases/phosphatases which open a new avenue for

4

the biotechnological production of mannitol.

Accepted Article

1

5 6

Results

7 8

MtlD of A. baylyi contains a phosphatase domain

9

10

A close inspection of the amino acid sequence of MtlD of A. baylyi revealed that the C-

11

terminus is similar to dehydrogenase domains found in other dehydrogenases with a

12

glycine-rich conserved domain (Rossmann-fold) starting at position 247 for cosubstrate

13

binding (GIHGFGAIGGG). Interestingly, the N-terminal domain of MtlD is similar to

14

members

15

phosphoesterases, ATPases, phosphatases, phosphonatases, dehalogenases and sugar

16

phosphomutases (Burroughs et al., 2006). The first 240 amino acids of MtlD share 43% of

17

the residues with a DL-glycerol-3-phosphatase (At5g57440) and 44% with another

18

phosphatase (At2g38740) of the HAD superfamily, present in Arabidopsis thaliana (Fig.

19

1). Sequence alignment of the phosphatase domain with the mannitol-1-phosphate

20

phosphatase (EsM1Pase) of Ectocarpus siliculosus, also a HAD-like phosphatase, showed

21

a similarity of 35%. All members of the HAD-family have a high sequence divergence but

22

share the α/β core domain catalytic scaffold that catalyzes the transfer of the phosphoryl

23

group. This core domain consists of four loops that contain highly conserved sequence

24

motifs (motif 1, DXD; motif 2, T/S; motif 3, K/R; motif 4, E/DD, GDXXXD or

25

GDXXXXD) (Caparros-Martin et al., 2013) (Fig. 1). Besides that core domain the

26

members of the HAD-family have a second, smaller cap domain which is used to divide

of

the

HAD

(haloacid

dehalogenase)

superfamily,

This article is protected by copyright. All rights reserved.

that

includes

6 the members into three subfamilies (Selengut, 2001). This second domain serves for

2

substrate recognition. Members of subfamily I have a small α-helical bundle cap domain

3

located between motif I and motif II, members of subfamily II have a mixed α/β domain

4

that is located between motif II and motif III and members of subfamily III have no cap

5

domain (Selengut, 2001; Lu et al., 2005). MtlD of A. baylyi has a large domain consisting

6

of α- helices between motif I and motif II which might serve as the cap and therefore we

7

conclude that MtlD belongs to subfamily I. The presence of this HAD-like domain

8

prompted us to look for a potential phosphatase activity of MtlD.

Accepted Article

1

9

10

Identification and biochemical characterization of the phosphatase activity of MtlD

11 12

To determine a potential phosphatase activity of MtlD we performed an enzyme assay with

13

pNPP (p-nitrophenylphosphate) as artificial substrate. These analyses revealed that purified

14

MtlD indeed dephosphorylates pNPP to pNP. Since phosphatases often require Mg2+ for

15

full activity (Zhang et al., 2004), Mg2+ dependence was analyzed. As can be seen from Fig.

16

2 phosphatase activity of MtlD was strongly dependent on the Mg2+ concentration and

17

maximal activity (103 mU/mg) was reached with 5 mM MgSO4. Activity was also

18

stimulated by other cations such as ZnCl2 (30%), CaCl2 (14%) or CuSO4 (15%) but Mg2+

19

stimulated the enzyme most effectively (MgCl2 100%; MgSO4 91%). Furthermore, there

20

was no phosphatase activity in the presence of the phosphatase inhibitor NaF (1 mM). The

21

pH optimum of the phosphatase reaction with pNPP as substrate was 5.0 but the enzyme

22

still retains an activity of 60% at pH 6.0 and 47% at pH 4.5 (Fig. S1).

23 24

MtlD is a bifunctional enzyme and exhibits Mtl-1-P phosphatase activity

25

This article is protected by copyright. All rights reserved.

7 31

To address the question whether MtlD is able to produce mannitol from F-6-P

2

was used as analytical technique to monitor online substrate consumption and product

3

formation. The pools of the phosphorylated metabolites, Mtl-1-P, Pi, and NADP+

4

produced, as well as F-6-P and NADPH consumption were monitored. The enzyme was

5

incubated in the absence of Mg2+ to prevent possible dephosphorylation of Mtl-1-P and the

6

reaction was started by addition of F-6-P at time zero (Fig. 3).

7

acquired sequentially for up to 50 minutes. From the array of spectra shown in Fig. 3 it is

8

apparent that the resonances due to F-6-P and NADPH decreased with concomitant

9

increase of the resonances corresponding to Mtl-1-P, Pi and NADP+. Control assays

10

without enzyme showed no release of Pi (Fig. S2). This confirms that MtlD catalyzes the

11

reduction of F-6-P to Mtl-1-P. Addition of 5 mM MgCl2 at time 21.3 min (indicated by an

12

arrow) restored phosphatase activity of MtlD and the peak due to Mtl-1-P decreased

13

immediately while Pi increased in the same proportion. The presence of mannitol as end-

14

product of the reaction was confirmed by 1H NMR (Fig. S3). These experiments clearly

15

prove that MtlD is a bifunctional enzyme possessing Mtl-1-P dehydrogenase activity and

16

Mg2+-dependent Mtl-1-P phosphatase activity.

Accepted Article

1

31

P-NMR

P-NMR spectra were

17 18

MtlD is highly specific for Mtl-1-P

19 20

To confirm that pure commercial Mtl-1-P is used as substrate by MtlD we analyzed the

21

dephosphorylation of Mtl-1-P in a similar assay monitored by 31P-NMR. In the presence of

22

MgCl2 (5 mM), Mtl-1-P was completely dephosphorylated by MtlD within 1.2 minutes of

23

reaction (Fig. 4A). In contrast, when the assay was repeated without addition of MgCl2, the

24

reaction proceeded at a much slower rate (Fig. 4B). The 1H-NMR spectra acquired at the

25

end of the assay revealed that in both cases Mtl-1-P was fully dephosphorylated to

This article is protected by copyright. All rights reserved.

8 mannitol. These results provide definite evidence for the presence of an Mg2+-dependent

2

phosphatase activity in MtlD. This phosphatase activity was highly specific for Mtl-1-P

3

and the activity determined was 83 U/mg. Substrates such as F-6-P, ribose-5-phosphate,

4

glucose-6-phosphate and glucose-1-phosphate were not dephosphorylated.

Accepted Article

1

5

The dehydrogenase activity of MtlD was previously found to increase with

6

increasing NaCl concentrations with highest activities at 500 mM NaCl (Sand et al.,

7

2013a). However still 40% of the maximal activity was observed with 300 mM NaCl. In

8

contrast, phosphatase activity (with pNPP as substrate) decreased with elevated NaCl

9

concentrations. However, in the presence of 300 mM NaCl still 20% of the maximal

10

activity were detected.

11 12

Distribution and phylogenetic analysis of bifunctional MtlD homologs

13 14

An analysis of around 8000 genomes retrieved from GenBank revealed that the

15

bifunctional MtlD enzyme is encoded in all 199 Acinetobacter strains except for six

16

Acinetobacter radioresistens strains. No other MtlD-like dehydrogenase was found in the

17

genomes of Acinetobacter. Genes encoding the bifunctional enzyme were also detected in

18

P. putida and a few other pseudomonads, and also in a few strains that belong to the

19

Clostridiales

20

Alicycliphilus and Methylobacillus as well as in Arcobacter, an ε-proteobacterium (Fig. 5).

21

In every case, the central region of the bifunctional enzyme contains the GXGXXG motif

22

that is typically found in the Rossmann fold and is responsible for binding to the

23

dinucleotide phosphate. Most of the strains that have the bifunctional MtlD encoded, have

24

no other MtlD-like domains encoded in their genomes. Phylogenetic analyses group all

(genera

Clostridium

and

Syntrophobotulus),

the

This article is protected by copyright. All rights reserved.

β-proteobacteria

9 Acinetobacter MtlDs in one cluster distant from homologs in other γ-proteobacteria and

2

homologs in other phyla.

Accepted Article

1

3 4

Discussion

5 6

Although mannitol is a major compatible solute in the soil bacterium A. baylyi, its

7

biosynthesis route remained elusive. Previous analysis revealed a pathway starting from

8

F-6-P via Mtl-1-P to mannitol but the phosphatase was not identified. The data presented

9

here provide clear evidence that MtlD catalyzes both steps of the pathway, the reduction of

10

F-6-P to Mtl-1-P and the subsequent dephosphorylation of Mtl-1-P to mannitol.

11

In general, Mtl-1-P dehydrogenases belong to the heterogenous group of long-chain

12

secondary alcohol dehydrogenases comprising 66 recognized members. The peptides of

13

this mainly prokaryotic protein family are typically between 350 and 560 amino acids long.

14

Mtl-1-P dehydrogenases (MtlDs) exhibit the shortest primary sequence within this protein

15

family (Kavanagh et al., 2003; Klimacek and Nidetzky, 2002). MtlD of A. baylyi

16

comprises 713 amino acids and thereby significantly extends the characteristic length of

17

secondary alcohol dehydrogenases. Moreover, only the conserved C-terminal domain, the

18

dehydrogenase part of MtlD, is similar to known MtlDs.

19

Sequence analysis revealed that A. baylyi MtlD contains indeed a phosphatase

20

domain at the N-terminus which could not be detected in an alignment with other known

21

MtlDs. This domain is conserved in members of the HAD superfamily which is a large

22

group of phosphatases with diverse substrate specifity (Burroughs et al., 2006; Caparros-

23

Martin et al., 2013). The sequence divergence of phosphatases in general is high but they

24

share an active site formed by four loops that accomodate the four consensus motifs

25

(motifs 1 - 4) (Morais et al., 2000; Lu et al., 2005). They were also detected at the N-

This article is protected by copyright. All rights reserved.

10 terminus of MtlD from A. baylyi. Loop 1 has an aspartate which acts as nucleophile and

2

mediates phosphoryl group transfer in HAD-like phosphatases. Loop 2 assists in substrate

3

binding and contributes threonine or serine and loop 3 has amino acids with positively

4

charged groups such as arginine or lysine and helps in orientation of the nucleophile or

5

electrophile (Caparros-Martin et al., 2013). Loop 4, together with loop 1, is essential for

6

cofactor binding and within the HAD superfamily of phosphatases there is a group of

7

magnesium dependent acid phosphatases which comprise a conserved aspartate box

8

(DXDX) near the N-terminus acting as the nucleophile residue (Collet et al., 1998)). In

9

these phosphatases, substrate binding is followed by closure of the active site. Interaction

10

of Mg2+ with the negatively charged phosphate is followed by a nucleophilic attack by the

11

first conserved aspartate (Caparros-Martin et al., 2013). Mg2+ which is essential for the

12

reaction, is bound to aspartate residues which are orientated in loop 4 that is also conserved

13

in MtlD of A. baylyi (Fig. 1).

Accepted Article

1

14

The phosphatase activity of the bifunctional MtlD is highly specific for Mtl-1-P.

15

This is consistent with the findings for other mannitol-1-phosphatases such as the

16

phosphatase from the red algae Caloglossa continua or from the brown algae

17

Spatoglossum pacificum and Dictyota dichotomoa that are also specific for Mtl-1-P (Ikawa

18

et al., 1972; Iwamoto et al., 2001; Kavanagh et al., 2003). Groisillier et al. (2014) also

19

described an HAD-like mannitol-1-phosphatase from Ectocarpus siliculosus that is also

20

highly specific for Mtl-1-P and they indicated that these enzymes constitute a new family

21

of phosphatases exhibiting a limited substrate specificity within the HAD superfamily that

22

usually show a broad substrate spectrum (Groisillier et al., 2014).

23

Mannitol is produced in response to osmotic stress and we have previously reported

24

that the Mtl-1-P dehydrogenase activity of MtlD increased with increasing salt

25

concentrations, with maximal activities at 500 mM NaCl (Sand et al., 2013a). In contrast,

This article is protected by copyright. All rights reserved.

11 the phosphatase activity decreased with increasing NaCl concentrations. This is consistent

2

with previous reports on the activity of other Mtl-1-P phosphatases such as the Mtl-1-P

3

phosphatase of E. siliculosus which decreased in the presence of 400 mM NaCl by 70% or

4

the Mtl-1-P phosphatase of C. continua which decreased at high salinity by 60% whereas

5

the dehydrogenase activity was stimulated (Iwamoto et al., 2001; Groisillier et al., 2014).

6

However, MtlD of A. baylyi exhibits both, dehydrogenase and phosphatase activities, in the

7

presence of moderate salt concentrations (300 mM NaCl) and even at lower salt

8

concentrations (< 300 mM NaCl) and in the absence of NaCl phosphatase activity was still

9

detected.

Accepted Article

1

10

It is interesting to note that bifunctional Mtl-1-P dehydrogenases/phosphatases are

11

widespread in members of the genus Acinetobacter but are mainly restricted to this genus.

12

A bioinformatics approach to identify candidate genes involved in carbon storage and

13

metabolism in the algae Micromonas spp. led to the detection of a potential Mtl-1-P

14

dehydrogenase gene encoding a HAD domain (Michel et al., 2010), although no

15

experimental evidence is available so far. Here, we provide clear experimental evidence

16

that MtlD of A. baylyi is a bifunctional enzyme mediating dehydrogenase and phosphatase

17

activity. Phylogenetic analyses however show that the resulting peptide is novel exclusive

18

of Acinetobacter and distant from orthologs in other taxa even within the same γ-

19

proteobacteria class. It is interesting to note that modules of MltD are more conserved

20

between proteins comprising of both modules than with others that contain just one, either

21

the phophatase or dehydrogenase module. Due to this finding it is tempting to speculate

22

that the two modules in the bifunctional phophatases/dehydrogenases have undergone

23

coevolution for an optimal cofunctioning of these different enzymatic activities.“

24

The fusion of two activites in one protein has taken place also in many other cases.

25

For instance, many pathways for di-myo-inositol-phosphate synthesis include bifunctional

This article is protected by copyright. All rights reserved.

12 enzymes while others include monofunctional proteins (Brito et al., 2011). Likewise, genes

2

involved in the biosynthetic pathways of amino acids such as tryptophan (HisB) and

3

histidine (AKAI) biosynthesis and others have been fused (Brilli and Fani, 2004; Fondi et

4

al., 2007).

Accepted Article

1

5

Gene duplication and fusion are two main mechanisms of the evolution of

6

metabolic pathways. Fusions serve for the physical association of different catalytic

7

domains or for generation of regulatory networks (Jensen, 1976). They often occur within

8

genes that code for proteins or enzymes that catalayze sequential steps within the same

9

metabolic pathway (Yanai et al., 2002). An advantage could be to simplify the regulation

10

of metabolic pathways and an enhanced catalytic activity due to channeling of

11

intermediates, such as Mtl-1-P, that are unstable or toxic for the cell (Fondi et al., 2009;

12

Nakano et al., 2013). The toxicity of sugar-phosphates in bacterial cells is well known. For

13

instances, an E. coli fructose-1-phosphate (F-1-P) kinase mutant which was unable to use

14

F-1-P or fructose as substrates was impaired in growth in the presence of these substrates

15

due to an accumulation of toxic F-1-P (Ferenci and Kornberg, 1973). Furthermore, an

16

MtlD-mutant of Salmonella typhimurium was also impaired in growth and lysed after the

17

addition of mannitol as nutrient. It is proposed that this effect was due to the accumulation

18

of Mtl-1-P which was toxic (Berkowitz, 1971).

19

The synthesis of compatible solutes such as mannitol occurs as response to high

20

salinity in the environment of a bacterial cell and has to be quite fast to prevent cell death.

21

As the cytoplasm of a cell is extremely complex and crowded this might be a barrier for

22

free diffusion and movement of proteins (Zimmerman and Trach, 1991; Dauty and

23

Verkman, 2004). Enzymes that are separated in the cytoplasm and have to interact are

24

therefore rate-limited in a fast interaction. The fusion of two polypeptides might support

25

substrate channeling and bypass this problem preventing the cell of wasting energy (Fani et

This article is protected by copyright. All rights reserved.

13 al., 2007). Combining the Mtl-1-P dehydrogenase and Mtl-1-P phosphatase domains in one

2

polypeptide might therefore lead to a more efficient reaction in the synthesis of mannitol in

3

response to osmotic stress.

Accepted Article

1

4

MtlD of A. baylyi represents a novel enzyme with a regulatory network that is

5

likewise expected to be previously unrecognized in the pathways leading to the synthesis

6

of mannitol. This could have implications in the design of chemicals that target

7

Acinetobacter as a pathogen or in biotechnological applications for the synthesis of

8

mannitol.

9

10

Experimental procedures

11 12

Enzyme purification and enzyme assays

13

MtlD (ACIAD1672) from A. baylyi ADP1 was heterologously expressed in E. coli grown

14

in LB medium and purified as described in Sand et al. (Sand et al., 2013a) except that for

15

the phosphatase assay with Mtl-1-P and other sugar substrates MOPS instead of sodium

16

phosphate buffer was used. A coupled enzymatic assay was used to determine phosphatase

17

activity of MtlD with pNPP as artificial substrate. A typical assay was carried out in a total

18

volume of 2 ml containing Na-acetate buffer (50 mM Na-acetate, 10 mM MgSO4), pH 5, at

19

37°C and 5 mM pNPP (p-nitrophenylphosphate). The amount of pNP released was

20

quantified by determining the absorbance at 405 nm.

21

Determination of the optimal pH was performed with pNPP in Na-acetate-MES-

22

HEPES buffer (each 50 mM). The determination of the activity in the presence of different

23

cations (5 mM) was performed in Na-acetate buffer (50 mM, pH 5.0) with pNPP as

24

substrate.

This article is protected by copyright. All rights reserved.

14 The phosphatase assays with Mtl-1-P and other sugar phosphates (each 0.5 mM) as

2

substrates were carried out by determining the amount of Pi released according to

3

Heinonen and Lahti (Heinonen and Lahti, 1981).

Accepted Article

1

4

Analysis of the substrate specificity was performed in 50 mM MOPS buffer (pH

5

7.0, 10 mM MgSO4) at 30°C. The enzyme was preincubated in the buffer for 10 min at

6

30°C before the reaction was started by addition of the substrates.

7 8

NMR analysis

9

31

P-NMR spectra were recorded at 202.45 MHz on a Bruker AVANCEII spectrometer

10

using a 5-mm selective probe head at 30ºC. Spectra were acquired with a 60º pulse and a

11

repetition delay of 1.2 seconds; each spectrum represents an average of 30 or 60 scans,

12

corresponding to 0.6 and 1.2 min acquisition, respectively. Proton broadband decoupling

13

was applied during the acquisition time only to avoid heating. 1H-NMR spectra were

14

acquired with the same spectrometer operating at 500.13 MHz using a 5-mm inverse

15

detection probe head. Water presaturation was applied.

16

Reactions were done in a total volume of 0.5 ml and assays with F-6-P as substrate

17

were prepared as follows: in a 5 mm NMR tube, 50 mM PIPES buffer (pH 6.5), 10 % (v/v)

18

D2O, 0.3 M NaCl, 10 mM NADPH, and 50 μg of MtlD, were combined and pre-incubated

19

at 30ºC before starting the reaction with the addition of 10 mM F-6-P. The reaction was

20

monitored online by acquiring successive

21

Mg2+, the sequence of spectra was interrupted, 5 mM MgCl2 was added to the same tube

22

and acquisition resumed under the same conditions until complete exustion of F-6-P. When

23

Mtl-1-P was studied as substrate the reaction mixture contained 50 mM PIPES buffer (pH

24

6.5), 10 % (v/v) D2O, 0.3 M NaCl, and 3.6 mM Mtl-1-P were combined in the presence or

25

absence of 5 mM MgCl2. The reaction was started by addition of 25 μg of MtlD. 3-

31

P-NMR spectra. To determine the effect of

This article is protected by copyright. All rights reserved.

15 (Trimethylsilyl)-1-propanesulfonic acid and 85% phosphoric acid contained in a cappillary

2

tube, both designated at 0 ppm were used as chemical shift references for 1H- and

3

NMR, respectively.

Accepted Article

1

31

P-

4 5

Phylogenetic analysis

6

The peptide files from available genomes in GenBank were downloaded and searched for

7

specific protein families using Pfam v25.0 run with HMMER3 (Eddy, 2008). A hit was

8

considered valid if its score was equal or bigger than the "gathering score" for the model.

9

Bifunctional MltD homologs were retrieved when the same peptide contained the domains

10

PF13419 (haloacid dehalogenase-like hydrolase that corresponds to the phosphatase

11

domain) and PF08125 (mannitol dehydrogenase, C-terminal). Analysis was carried out by

12

customized Python scripts.

13

The alignments to generate the maximum likelihood trees were generated using

14

MUSCLE (Edgar, 2004) and trimmed using the trimAl software (Capella-Gutierrez et al.,

15

2009) to eliminate highly diverged regions. The maximum-likelihood tree was inferred

16

with RAxML (Stamatakis, 2006) using the LG model with gamma distribution of rates and

17

invariant site categories (implemented as “PROTGAMMAILG”). To obtain the statistical

18

confidence of branching order, 100 pseudoreplicates were generated. The results with

19

maximum-likelihood were confirmed by Bayesian inference using MrBayes v. 3.2.2

20

(Ronquist and Huelsenbeck, 2003). Two independent 1,000,000 generations were run with

21

the WAG model and trees were sampled every 500 generations. The average standard

22

deviation of split frequencies was below 0.007 at the end, which indicates that the MCMC

23

chains converged. The Bayesian tree of the bifunctional MltD amino acid sequences

24

constructed yielded a branching order similar to the maximum likelihood tree.

25

This article is protected by copyright. All rights reserved.

16 Acknowledgments

Accepted Article

1 2 3

This work was funded by a grant from the Deutsche Forschungsgemeinschaft. JMG was

4

funded

5

MarineGems (CTM2010-20361) from the Spanish Ministry of Science and Innovation.

6

The NMR spectrometers are part of The National NMR Facility, supported by Fundação

7

para a Ciência e a Tecnologia (RECI/BBB-BQB/0230/2012).

by

the

CONSOLIDER-INGENIO2010

Program

(CSD2008-00077)

and

8 9

References

10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39

Berkowitz, D. (1971) D-Mannitol utilization in Salmonella typhimurium. J Bacteriol 105: 232-240. Brilli, M., and Fani, R. (2004) Molecular evolution of hisB genes. J Mol Evol 58: 225-237. Brito, J.A., Borges, N., Vonrhein, C., Santos, H., and Archer, M. (2011) Crystal structure of Archaeoglobus fulgidus CTP:inositol-1-phosphate cytidylyltransferase, a key enzyme for di-myo-inositol-phosphate synthesis in (hyper)thermophiles. J Bacteriol 193: 2177-2185. Burroughs, A.M., Allen, K.N., Dunaway-Mariano, D., and Aravind, L. (2006) Evolutionary genomics of the HAD superfamily: understanding the structural adaptations and catalytic diversity in a superfamily of phosphoesterases and allied enzymes. J Mol Biol 361: 1003-1034. Caparros-Martin, J.A., McCarthy-Suarez, I., and Culianez-Macia, F.A. (2013) HAD hydrolase function unveiled by substrate screening: enzymatic characterization of Arabidopsis thaliana subclass I phosphosugar phosphatase AtSgpp. Planta 237: 943954. Capella-Gutierrez, S., Silla-Martinez, J.M., and Gabaldon, T. (2009) trimAl: a tool for automated alignment trimming in large-scale phylogenetic analyses. Bioinformatics 25: 1972-1973. Collet, J.F., Stroobant, V., Pirard, M., Delpierre, G., and Van Schaftingen, E. (1998) A new class of phosphotransferases phosphorylated on an aspartate residue in an aminoterminal DXDX(T/V) motif. J Biol Chem 273: 14107-14112. da Costa, M.S., Santos, H., and Galinski, E.A. (1998) An overview of the role and diversity of compatible solutes in Bacteria and Archaea. Adv Biochem Eng Biotechnol 61: 117-153. Dauty, E., and Verkman, A.S. (2004) Molecular crowding reduces to a similar extent the diffusion of small solutes and macromolecules: measurement by fluorescence correlation spectroscopy. J Mol Recogni 17: 441-447. Eddy, S.R. (2008) A probabilistic model of local sequence alignment that simplifies statistical significance estimation. PLoS Comput Biol 4: e1000069.

This article is protected by copyright. All rights reserved.

17 Edgar, R.C. (2004) MUSCLE: multiple sequence alignment with high accuracy and high throughput. Nucleic Acids Res 32: 1792-1797. Fani, R., Brilli, M., Fondi, M., and Lio, P. (2007) The role of gene fusions in the evolution of metabolic pathways: the histidine biosynthesis case. BMC Evol Biol 7 Suppl 2: S4. Ferenci, T., and Kornberg, H.L. (1973) The utilization of fructose by Escherichia coli. Properties of a mutant defective in fructose 1-phosphate kinase activity. Biochem J 132: 341-347. Fondi, M., Brilli, M., and Fani, R. (2007) On the origin and evolution of biosynthetic pathways: integrating microarray data with structure and organization of the common pathway genes. BMC Bioinformatics 8 Suppl 1: S12. Fondi, M., Emiliani, G., and Fani, R. (2009) Origin and evolution of operons and metabolic pathways. Res Microbiol 160: 502-512. Galinski, E.A., and Trüper, H.G. (1994) Microbial behaviour in salt-stressed ecosystems. FEMS Microbiol Rev 15: 95-108. Gaspar, P., Neves, A.R., Gasson, M.J., Shearman, C.A., and Santos, H. (2011) High yields of 2,3-butanediol and mannitol in Lactococcus lactis through engineering of NAD+ cofactor recycling. Appl Environ Microbiol 77: 6826-6835. Groisillier, A., Shao, Z., Michel, G., Goulitquer, S., Bonin, P., Krahulec, S. et al. (2014) Mannitol metabolism in brown algae involves a new phosphatase family. J Exp Bot 65: 559-570. Heinonen, J.K., and Lahti, R.J. (1981) A new and convenient colorimetric determination of inorganic orthophosphate and its application to the assay of inorganic pyrophosphatase. Anal Biochem 113: 313-317. Ikawa, T., Watanabe, T., and Nisizawa, K. (1972) Enzymes involved in the last steps of the biosynthesis of mannitol in brown algae. Plant and Cell Physiol 13: 1017-1029. Iwamoto, K., Kawanobe, H., Shiraiwa, Y., and Ikawa, T. (2001) Purification and characterization of mannitol-l-phosphatase in the red alga Caloglossa continua (Ceramiales, Rhodophyta). Mar Biotechnol 3: 493-500. Iwamoto, K., Kawanobe, H., Ikawa, T., and Shiraiwa, Y. (2003) Characterization of saltregulated mannitol-1-phosphate dehydrogenase in the red alga Caloglossa continua. Plant Physiol 133: 893-900. Jennings, D.H. (1984) Polyol metabolism in fungi. Adv Microb Physiol 25: 149-193. Jensen, R.A. (1976) Enzyme recruitment in evolution of new function. Annu Rev Microbiol 30: 409-425. Kavanagh, K.L., Klimacek, M., Nidetzky, B., and Wilson, D.K. (2003) Crystal structure of Pseudomonas fluorescens mannitol 2-dehydrogenase: evidence for a very divergent long-chain dehydrogenase family. Chem Biol Interact 143-144: 551-558. Kets, E.P., Galinski, E.A., de Wit, M., de Bont, J.A., and Heipieper, H.J. (1996) Mannitol, a novel bacterial compatible solute in Pseudomonas putida S12. J Bacteriol 178: 6665-6670. Klimacek, M., and Nidetzky, B. (2002) A catalytic consensus motif for D-mannitol 2dehydrogenase, a member of a polyol-specific long-chain dehydrogenase family, revealed by kinetic characterization of site-directed mutants of the enzyme from Pseudomonas fluorescens. Biochem J 367: 13-18. Loescher, W.H., Tyson, R.H., Everard, J.D., Redgwell, R.J., and Bieleski, R.L. (1992) Mannitol synthesis in higher plants: evidence for the role and characterization of a NADPH-dependent mannose 6-phosphate reductase. Plant Physiol 98: 1396-1402. Lu, Z., Dunaway-Mariano, D., and Allen, K.N. (2005) HAD superfamily phosphotransferase substrate diversification: structure and function analysis of HAD subclass IIB sugar phosphatase BT4131. Biochemistry 44: 8684-8696.

Accepted Article

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48 49 50

This article is protected by copyright. All rights reserved.

18 Michel, G., Tonon, T., Scornet, D., Cock, J.M., Kloareg, B. (2010) Central and storage carbon metabolism of the brown alga Ectocarpus siliculosus: insights into the origin and evolution of storage carbohydrates in Eukaryotes. New Phytol 188: 67-81. Morais, M.C., Zhang, W., Baker, A.S., Zhang, G., Dunaway-Mariano, D., and Allen, K.N. (2000) The crystal structure of Bacillus cereus phosphonoacetaldehyde hydrolase: insight into catalysis of phosphorus bond cleavage and catalytic diversification within the HAD enzyme superfamily. Biochemistry 39: 10385-10396. Nakano, T., Ohki, I., Yokota, A., and Ashida, H. (2013) MtnBD is a multifunctional fusion enzyme in the methionine salvage pathway of Tetrahymena thermophila. PloS one 8: e67385. Neves, A.R., Ramos, A., Shearman, C., Gasson, M.J., Almeida, J.S., and Santos, H. (2000) Metabolic characterization of Lactococcus lactis deficient in lactate dehydrogenase using in vivo 13C-NMR. Eur J Biochem 267: 3859-3868. Roeßler, M., and Müller, V. (2001) Osmoadaptation in bacteria and archaea: common principles and differences. Environ Microbiol 3: 743-754. Ronquist, F., and Huelsenbeck, J.P. (2003) MrBayes 3: Bayesian phylogenetic inference under mixed models. Bioinformatics 19: 1572-1574. Rumpho, M.E., Edwards, G.E., and Loescher, W.H. (1983) A pathway for photosynthetic carbon flow to mannitol in celery leaves: activity and localization of key enzymes. Plant Physiol 73: 869-873. Sand, M., Mingote, A.I., Santos, H., Müller, V., and Averhoff, B. (2013a) Mannitol, a compatible solute synthesized by Acinetobacter baylyi in a two-step mechanism including a salt-induced and salt-dependent mannitol-1-phosphate dehydrogenase. Environ Microbiol 15: 2187-2197. Sand, M., Stahl, J., Waclawska, I., Ziegler, C., and Averhoff, B. (2013b) Identification of an osmo-dependent and an osmo-independent choline transporter in Acinetobacter baylyi: implications in osmostress protection and metabolic adaptation. Environ Microbiol doi: 10.1111/1462-2920.12188 Sand, M., de Berardinis, V., Mingote, A., Santos, H., Göttig, S., Müller, V., and Averhoff, B. (2011) Salt adaptation in Acinetobacter baylyi: identification and characterization of a secondary glycine betaine transporter. Arch Microbiol 193: 723-730. Selengut, J.D. (2001) MDP-1 is a new and distinct member of the haloacid dehalogenase family of aspartate-dependent phosphohydrolases. Biochemistry 40: 12704-12711. Stamatakis, A. (2006) RAxML-VI-HPC: maximum likelihood-based phylogenetic analyses with thousands of taxa and mixed models. Bioinformatics 22: 2688-2690. Stoop, M.H., Williamson, J.D., and Pharr, D.M. (1996) Mannitol metabolism in plants: a method for coping with stress. Trends Plant Sci 1: 139-144. Wisselink, H.W., Weusthuis, R.A., Eggink, G., Hugenholtz, J., and Grobben, G.J. (2002) Mannitol production by lactic acid bacteria: a review. Int Dairy J 12: 151-161. Wood, J.M., Bremer, E., Csonka, L.N., Krämer, R., Poolman, B., van der Heide, T., and Smith, L.T. (2001) Osmosensing and osmoregulatory compatible solute accumulation by bacteria. Comp Biochem Physiol A Mol Integr Physiol 130: 437-460. Yanai, I., Wolf, Y.I., and Koonin, E.V. (2002) Evolution of gene fusions: horizontal transfer versus independent events. Genome Biol 3: research0024. Zhang, G., Morais, M.C., Dai, J., Zhang, W., Dunaway-Mariano, D., and Allen, K.N. (2004) Investigation of metal ion binding in phosphonoacetaldehyde hydrolase identifies sequence markers for metal-activated enzymes of the HAD enzyme superfamily. Biochemistry 43: 4990-4997.

Accepted Article

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20 21 22 23 24 25 26 27 28 29 30 31 32 33 34 35 36 37 38 39 40 41 42 43 44 45 46 47 48

This article is protected by copyright. All rights reserved.

19 Zimmerman, S.B., and Trach, S.O. (1991) Estimation of macromolecule concentrations and excluded volume effects for the cytoplasm of Escherichia coli. J Mol Biol 222: 599-620.

Accepted Article

1 2 3 4 5 6

Figure legends

7 8

Fig. 1. Conserved domains of the mannitol-1-phosphate dehydrogenase MtlD of A. baylyi.

9

MtlD has a HAD-like domain at the N-terminus and a mannitol-1-phosphate

10

dehydrogenase-domain (MtlD) and the Rossmann-fold at the C-terminus (A). The amino

11

acid sequence of the phosphatase domain of MtlD of A. baylyi, the two phosphatases of A.

12

thaliana At5g57440 and At2g38740 and the mannitol-1-phosphate phosphatase of

13

Ectocarpus siliculosus EsM1Pase were aligned with the program ClustalW. Residues that

14

are identical or similar in At5g57440, At2g38740 and EsM1Pase are highlighted with

15

asterisks or dots, respectively, and shared motifs are highlighted in grey (B).

16 17

Fig. 2. Mg2+ dependence of MtlD activity. The assay was performed in 50 mM Na-acetate

18

buffer, pH 5.0, with 20 µg enzyme which was preincubated for 10 minutes at 37°C in

19

buffer with the MgSO4 concentrations as indicated. The reaction was started by addition of

20

pNPP to a final concentration of 5 mM. At different time points, 200 µl samples were

21

taken and the reaction was stopped with 800 µl 2 M NaOH. The release of pNP was

22

measured at 405 nm. The figure shows one of three replicates.

23

31

24

Fig. 3. Sequence of

P-NMR spectra monitoring the reaction products of MtID activity.

25

The reaction mixture contained 50 mM PIPES, pH 6.5, 0.3 M NaCl, 10 mM F-6-P, 10 mM

26

NADPH and 50 μg MtID. The reaction was performed at 30ºC. F-6-P was added at time

27

point zero and each spectrum was acquired for 1.2 min, starting at the times indicated. At

28

time 21.3 min, 5 mM MgCl2 (indicated by the arrow) was added. The shift in the

This article is protected by copyright. All rights reserved.

20 resonances is due to pH changes due to insufficient buffer capacity. ●, Pi; ■, Mtl-1-P; ▲,

2

NADP+; △, F6P; ☐, NADPH.

Accepted Article

1

3 4

Fig. 4. Dephosphorylation of Mtl-1-P by MtlD followed by 31P-NMR. Sequences of 31P-

5

NMR acquired consecutively after addition of MtID (25 μg) in the presence of 5 mM

6

MgCl2 (A) or in the absence of MgCl2 (B). Reaction mixture contained 50 mM PIPES, pH

7

6.5, 0.3 M NaCl, 25 μg MtID, and 3.6 mM Mtl-1-P. MtlD was added at time point zero.

8

Each spectrum was acquired for 0.6 min, starting at the time points indicated. ●, Pi; ■, Mtl-

9

1-P.

10 11

Fig. 5. Maximum-likelihood phylogenetic tree of representative bifunctional MltD amino

12

acid sequences. The tree was constructed using RAxML v. 7.8.6. Acinetobacter peptides

13

cluster away from the rest of bifunctional MltD homologs. This topology was found with

14

both Bayesian inference and maximum likelihood methods. Numbers at nodes are

15

bootstrap values in 100 replicates. The scale bar indicates substitutions per site.

This article is protected by copyright. All rights reserved.

Accepted Article

21

1 2

emi_12503_f1

3

This article is protected by copyright. All rights reserved.

22

Accepted Article

1

2 3

emi_12503_f2

4

This article is protected by copyright. All rights reserved.

23

Accepted Article

1

2 3

emi_12503_f3

4

This article is protected by copyright. All rights reserved.

24

Accepted Article

1

2 3

emi_12503_f4

4

This article is protected by copyright. All rights reserved.

25

Accepted Article

1

2 3

emi_12503_f5

4

This article is protected by copyright. All rights reserved.

phosphatases: a family of novel bifunctional enzymes for bacterial adaptation to osmotic stress.

The nutritionally versatile soil bacterium Acinetobacter baylyi ADP1 copes with salt stress by the accumulation of compatible solutes, a strategy that...
1MB Sizes 3 Downloads 3 Views