31
1 Functional Characterization of 2-Hydroxy-1-naphthoic Acid Non-oxidative 1 Decarboxylase from Burkholderia sp. Strain BC1: A Novel Member of Amidohydrolase- 2 2 Protein Family 3 4 5 Piyali Pal Chowdhury, Soumik Basu, Arindam Dutta and Tapan K. Dutta# 6 Department of Microbiology, Bose Institute, Kolkata, India 7 8 Running Title: Non-oxidative decarboxylase from Burkholderia sp. 9 10 11 #Address correspondence to Tapan K. Dutta, [email protected] 12 13 14 15 16 17 This manuscript contains supplementary data 18 19 20 21 22 23 JB Accepted Manuscript Posted Online 11 April 2016 J. Bacteriol. doi:10.1128/JB.00250-16 Copyright © 2016, American Society for Microbiology. All Rights Reserved. on April 16, 2018 by guest http://jb.asm.org/ Downloaded from

Functional Characterization of 2-Hydroxy-1-naphthoic Acid Non

  • Upload
    vonhu

  • View
    221

  • Download
    2

Embed Size (px)

Citation preview

1

Functional Characterization of 2-Hydroxy-1-naphthoic Acid Non-oxidative 1

Decarboxylase from Burkholderia sp. Strain BC1: A Novel Member of Amidohydrolase-2

2 Protein Family 3

4

5

Piyali Pal Chowdhury, Soumik Basu, Arindam Dutta and Tapan K. Dutta# 6

Department of Microbiology, Bose Institute, Kolkata, India 7

8

Running Title: Non-oxidative decarboxylase from Burkholderia sp. 9

10

11

#Address correspondence to Tapan K. Dutta, [email protected] 12

13

14

15

16

17

This manuscript contains supplementary data 18

19

20

21

22

23

JB Accepted Manuscript Posted Online 11 April 2016J. Bacteriol. doi:10.1128/JB.00250-16Copyright © 2016, American Society for Microbiology. All Rights Reserved.

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

2

ABSTRACT 24

The gene encoding a non-oxidative decarboxylase capable of catalyzing the transformation of 25

2-hydroxy-1-naphthoic acid (2H1NA) to 2-naphthol was identified, recombinantly expressed 26

and purified to homogeneity. The putative gene sequence of the decarboxylase (hndA) 27

encodes a 316 amino acid protein (HndA) with a predicted molecular mass of 34 kDa. HndA 28

exhibited high identity with uncharacterized amidohydrolase-2 proteins of various 29

Burkholderia sp. whereas it showed a modest 27% identity with γ-resorcylate decarboxylase, 30

a well characterized non-oxidative decarboxylase belonging to the amidohydrolase 31

superfamily. Biochemically characterized HndA demonstrated strict substrate specificity 32

towards 2H1NA while inhibition studies with HndA indicated the presence of zinc as the 33

transition metal centre as confirmed by atomic absorption spectroscopy. A three-dimensional 34

structural model of HndA followed by docking analysis identified the conserved metal 35

coordinating and substrate binding residues, while their importance in catalysis was validated 36

by site-directed mutagenesis. 37

38

IMPORTANCE 39

Microbial non-oxidative decarboxylases play a crucial role in the metabolism of a large array 40

of carboxy aromatic chemicals released into environment from a variety of natural and 41

anthropogenic sources. Among these, hydroxynaphthoic acids are usually encountered as 42

pathway intermediates in the bacterial degradation of polycyclic aromatic hydrocarbons. The 43

present study reveals biochemical and molecular characterization of a 2-hydroxy-1-naphthoic 44

acid non-oxidative decarboxylase involved in an alternative metabolic pathway which can be 45

classified as a member of the small repertoire of non-oxidative decarboxylases belonging to 46

the amidohydrolase-2 family of proteins. The strict substrate specificity and sequence 47

uniqueness makes it a novel member of the metallo-dependent hydrolase superfamily. 48

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

3

49

INTRODUCTION 50

Decarboxylase is one of the most important classes of enzymes involved in a large variety of 51

catabolic and anabolic pathways. The majority of the decarboxylases utilize an organic 52

cofactor or a transition metal coupled with dioxygen to activate their substrates leading to the 53

removal of carbon dioxide (1). However, there is a small group of transition metal-dependent 54

decarboxylases that carry out decarboxylation of various aromatic acids in a non-oxidative 55

manner. These non-oxidative decarboxylases act on various lignin-derived compounds such 56

as 4-hydroxybenzoic acid (2) (carboxy)vanillic acid (3, 4), protocatechuic acid (5), ferulic 57

acid (6), p-coumaric acid (7) and estrogenic phthalate (8). Likewise, oxygen-independent 58

decarboxylases are also involved in 2-nitrobenzoic acid degradation pathway (9, 10), 59

tryptophan catabolic pathway (11) and thymidine salvage pathway (12). 60

Non-oxidative decarboxylases, in general, can broadly be classified into two major 61

groups depending on their oxygen sensitivity. Oxygen-sensitive decarboxylases, viz. 4-62

hydroxybenzoate decarboxylase (2), 3,4-dihydroxybenzoate decarboxylase (5) and indole-3-63

carboxylate decarboxylase (13) catalyze reversible reactions, both carboxylation and 64

decarboxylation. On the other hand, oxygen-insensitive decarboxylases such as 2,3-65

dihydroxybenzoate decarboxylase (14), 5-carboxyvanillate decarboxylase (4) and 4,5-66

dihydroxyphthalate decarboxylase (8) have been reported to catalyze the decarboxylation 67

reaction only. However, there are a few non-oxidative oxygen-insensitive decarboxylases, 68

viz. γ-resorcylate decarboxylase (15), vanillate/4-hydroxybenzoate decarboxylase (16) and 69

salicylate decarboxylase (17) that have been documented to catalyze reversible reactions. 70

Hydroxynaphthoates, such as 1-hydroxy-2-naphthoic acid (1H2NA), 2-hydroxy-1-71

naphthoic acid (2H1NA) and 3-hydroxy-2-naphthoic acid (3H2NA) are normally 72

encountered during the bacterial degradation of polycyclic aromatic hydrocarbons (PAHs), 73

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

4

viz. phenanthrene, anthracene and pyrene, and are metabolized either through ring cleavage 74

(18-22) or by oxidative decarboxylation (23). In addition, decarboxylation of 1H2NA to 1-75

naphthol has been proposed, based on the identification of the later compound during 76

degradation of phenanthrene in a few bacteria (24-26). Similarly, decarboxylation of 77

phenanthrene-4,5-dicarboxylic acid to phenanthere-4-carboxylic acid has also been reported 78

in the pathway of pyrene degradation (19). However there is no documented report of any 79

non-oxidatively decarboxylated product, produced during the metabolism of 2H1NA or 80

3H2NA. Despite published reports of purification and characterization of several non-81

oxidative hydroxybenzoate decarboxylases (5, 14-16), there are no examples of any enzyme 82

catalyzing non-oxidative decarboxylation of any of the hydroxynaphthoic acid isomers. 83

Previously, we had reported a non-conventional degradation pathway of 2H1NA in 84

Burkholderia sp. BC1 describing 2-naphthol, gentisaldehyde and gentisic acid as pathway 85

intermediates (27). Moreover, presence of a strictly inducible non-oxidative decarboxylase 86

was also observed in the cell-free extract of 2H1NA-grown culture catalyzing the enzymatic 87

transformation of 2H1NA to 2-naphthol. In the present study, we describe a proteomic 88

approach-based gene cloning and functional characterization of non-oxidative 2H1NA 89

decarboxylase from Burkholderia sp. BC1. In addition, the roles of specific amino acid 90

residues responsible for substrate binding and enzyme catalysis have been elucidated. 91

92

MATERIALS AND METHODS 93

Bacterial strains, plasmids and culture conditions. Strains and plasmids used in this study 94

are listed in Table S1 in supplemental material. Recombinant constructs in E. coli [XL1-Blue 95

and BL21(DE3)] were routinely grown and maintained in Luria-Bertani (LB) broth (per litre) 96

containing 10 g bactotryptone, 5 g yeast extract, and 10 g NaCl; pH 7.2 or on LB solid 97

medium (1.8 % w/v agar) at 37°C. Where appropriate, ampicillin (100 µg/ml), kanamycin 98

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

5

(50 µg/ml), chloramphenicol (12.5 µg/ml), isopropyl-β-D-thiogalactopyranoside (IPTG) 99

(0.1 to 1 mM) or 5‐bromo‐4‐chloro‐3‐indolyl β‐D‐galactopyranoside (X‐gal) (20 µg/ml) 100

was added. For expression cloning, pET28a (Novagen, Madison, WI) served as the 101

expression vector. 102

Partial purification and gene identification of 2H1NA decarboxylase. Native 2H1NA 103

decarboxylase was purified from crude cell-free extract of BC1 cells grown for 16 h at 28°C 104

in four liters of MSM (20) containing 0.5 g/liter 2H1NA. Crude cell-free extract was prepared 105

as described previously (27), which was then fractionated by sequential protein precipitation 106

using ammonium sulphate. The 30-50% ammonium sulphate saturated fraction was 107

centrifuged at 12000×g for 30 min and the resulting pellet was dissolved in buffer A (50 mM 108

K2HPO4-KH2PO4 buffer, pH 7.0) and dialyzed against buffer B (50 mM K2HPO4-KH2PO4 109

buffer, pH 7.0 containing 0.8 M (NH4)2SO4). The dialyzed fraction was then loaded onto a 110

column (2.5 cm×10 cm), packed with Phenyl Sepharose 6 Fast Flow, pre-equilibrated with 111

buffer B. The column was washed with five column volumes of buffer B and then the 112

adsorbed proteins were eluted in steps using 10 column volumes of each of buffer A 113

containing different concentration of (NH4)2SO4 (0.8 ̶ 0.05 M). Finally, the column was 114

washed with two column volumes of buffer A. All purification steps were carried out at 4°C 115

or on ice under aerobic conditions. Fractions exhibiting 2H1NA decarboxylase activity were 116

combined and dialyzed against buffer C (50 mM K2HPO4-KH2PO4 buffer, pH 7.0, 10% 117

glycerol), concentrated by ultrafiltration (Millipore, Massachusetts, USA) and stored at -80°C 118

until further use. The purity of the protein fractions obtained after ammonium sulphate 119

precipitation and hydrophobic interaction chromatography was evaluated by 12.5% SDS-120

PAGE analysis followed by Coomassie blue staining in the presence of pre-stained protein 121

molecular mass markers (Puregene, Genetix, India) by standard techniques. Protein 122

quantification was done using the method of Bradford (28). 123

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

6

For identification of the decarboxylase, tryptic digestion of the protein and subsequent 124

extraction of peptides from SDS-PAGE gel matrices were carried out by methods described 125

by Shevchenko et al. (29) followed by MALDI-TOF MS and MS/MS analyses using 126

AutoFlex II (Bruker Daltonics, Germany) MALDI-tandem time-of-flight (TOF/TOF) mass 127

spectrometer equipped with a pulsed N2 laser (λ = 337 nm, 50 Hz). The mass spectra were 128

analysed with Flex Analysis Software (version 2.4, Bruker, Daltonics). From the MS/MS 129

data, partial amino acid sequences of the peptides were determined using PEAKS studio 7 130

(Bioinformatics solutions, Ontario, Canada) and the peptide sequences were subjected to 131

blastp (30) analysis for identification. Primers [HNDA_F, 5′-TGCTGTCGCTGACGGC-3′ and 132

HNDA_R, 5′-TTGCTGAGCAGCACGAC-3′] were designed on the basis of conserved regions 133

exhibited in multiple sequence alignment, generated by Clustalx v1.81 (31) using 134

amidohydrolase gene sequences of various Burkholderia sp. (see Table S2 in supplemental 135

material). Using the primers, PCR was carried out in a 50-µl reaction volume using phusion 136

DNA polymerase (Thermo Fischer) in a MJ Mini Gradient Thermal Cycler (Bio-Rad 137

Laboratories, Inc., Hercules, CA, USA) with the following thermo-cycling conditions: 30 s at 138

98°C followed by 30 cycles of 30 s at 98°C, 30 s at 55°C and 10 s at 72°C. Final extension 139

was performed at 72°C for 7:00 min. The resulting PCR product was sequenced as reported 140

previously (27) 141

Enzyme assay. Non-oxidative decarboxylase activity was qualitatively determined by UV-142

visible spectral analysis as described previously (27) while the activity was quantitatively 143

determined based on the formation of 2-naphthol, analyzed by HPLC using a methanol/water 144

(50:50 v/v) isocratic solvent system with a flow rate of 1 ml/min (27). A standard curve of 2-145

naphthol created by HPLC under identical analytical conditions was used for quantitative 146

estimation. One unit of enzyme activity is defined as the amount of enzyme required for the 147

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

7

production of 1 μmol of product per min. Specific activity is expressed as units per mg of 148

protein. 149

For measurement of carboxylase activity, a standard reaction mixture containing 150

recombinant protein (100 μg), 2-naphthol (20 mM) and NaHCO3/NH4HCO3 (1.0 or 2.5 M) in 151

a final volume of 1 ml of buffer A was prepared and the reaction mixture was incubated for 152

60 min at 35°C. To analyze the reaction product, HPLC analysis was performed as described 153

above. 154

Fosmid library construction, screening and sequence analysis of amidohydrolase gene. 155

A genomic library of strain BC1 was prepared in E. coli using CopyControl™ HTP Fosmid 156

Library Production Kit (Epicentre, Madison, Wisconsin) according to the manufacturer’s 157

protocol (Epicentre, Madison, Wisconsin). The resulting fosmid library was screened by 158

PCR using HNDA_F and HNDA_R primers for clones harbouring 2H1NA decarboxylase 159

gene as described above. Fosmid DNA was isolated from the PCR positive fosmid clones 160

using the FosmidMAX™ DNA Purification Kit (Epicentre, Madison, Wisconsin), digested 161

with EcoRI, HindIII and SacII enzymes and the DNA fragments (1-8 kb) were subcloned in 162

pBluescript SK(-) vector. The colonies were rescreened by PCR using the same primer pair. 163

The recombinant plasmids from the screened colonies were individually isolated and 164

sequenced using M13 universal sequencing primers. The sequences were analyzed by 165

BLAST analysis (version 2.2.12, National Center for Biotechnology Information) and the 166

gaps between genes were bridged by conventional primer walking method. 167

Cloning, expression and purification of recombinant proteins. Primers, Ex_HNDA_F [5′-168

CCGGAATTCATGACCGACCATCACCGTATC-3′] and Ex_HNDA_R [5′-169

CCCAAGCTTTTGTTGTGTTGTTGCGTCAG-3′] were designed (restriction endonuclease 170

recognition sites are underlined for EcoRI and HinDIII respectively) to amplify the complete 171

2H1NA decarboxylase gene (hndA) from genomic DNA of strain BC1. The amplified PCR 172

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

8

product was digested with EcoRI and HinDIII and ligated into similarly digested pET28a 173

expression vector to form pET28a:HndA. The resulting plasmid was transformed into E. coli 174

BL21(DE3) and plated on LB-agar plates containing kanamycin. For the preparation of single 175

amino acid substitution in HndA, pET28:HndA was subjected to whole plasmid PCR with 176

mutagenic primers (see Table S3 in Supplemental material) under following thermo-cycling 177

conditions: 3 min at 98°C, followed by 16 cycles of 30 s at 98°C, 30 s at 55°C and 3 min at 178

72°C with final extension at 72°C for 10 min using phusion high-fidelity DNA polymerase 179

(Thermo Fischer Scientific). After digestion with DpnI for 2 h, the PCR product was 180

transformed into electrocompetant E. coli Top10 and plated on LB-agar plates containing 181

kanamycin. Plasmids isolated from random clones were subjected to sequencing analysis to 182

confirm the mutation at specified location. 183

For recombinant enzyme expression and purification, E. coli BL21(DE3) cells harbouring 184

pET28a:HndA or its mutant derivatives (see Table S1 in Supplemental material) were grown 185

in 500 ml LB medium at 37°C with kanamycin to achieve an OD600 of 0.5 followed by the 186

addition of 0.5 mM IPTG (final concentration) and grown further at 28°C for 3 h. The 187

cultures were harvested by centrifugation (8000×g) and lysed in 10 ml of lysis buffer (50 mM 188

NaHPO4, 300 mM NaCl, 10 mM imidazole and 10 % glycerol) using a pre-cooled French 189

press (Constant Cell Disruption System, One Shot model, United Kingdom) at 18 Kpsi for 190

one cycle. After removal of the cell debris, the supernatant containing the His6-tagged wild 191

type or mutant recombinant protein was purified by Ni2+-NTA agarose affinity 192

chromatography using the purification buffers [wash buffer (50 mM NaHPO4, 300 mM NaCl, 193

40 mM imidazole and 10% glycerol) and elution buffer (50 mM NaHPO4, 300 mM NaCl, 194

250 mM imidazole and 10% glycerol)] according to the manufacturer’s instructions (Qiagen). 195

The purified protein fractions were pooled, dialyzed against buffer C and analyzed by 12.5% 196

SDS-PAGE. The dialyzed protein preparation was used in all biochemical studies. 197

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

9

Phylogenetic analysis. The amino acid sequences of various proteins belonging to the 198

amidohydrolase-1 and amidohydrolase-2 families were retrieved from NCBI (see Table S4 in 199

Supplemental material), aligned and a phylogenetic tree was constructed using the neighbour-200

joining algorithm as implemented in ClustalX v1.81 (31). The tree was visualized using the 201

program Tree Explorer v2.12, a stand-alone version of the same program implemented in 202

MEGA 5 (32). 203

Gel filtration. Native molecular mass of the decarboxylase was estimated by gel filtration 204

chromatography using a P4000 PolySep GFC column (30×0.7 cm, Phenomenax, Torrance, 205

CA), equilibrated with buffer A containing 200 mM NaCl. Flow rate used was 0.5 ml/min. 206

Yeast alcohol dehydrogenase (150 kDa), conalbumin (75 kDa), ovalbumin (44 kDa), 207

carbonic anhydrase (29 kDa) and ribonuclease A (13.7 kDa) were used as standard proteins. 208

Blue dextran (2000 kDa) was used to calculate the void volume. 209

Biochemical studies. All the enzyme kinetic analyses were done at 35°C and pH 7.5. For 210

decarboxylation, the kinetic data were assessed using 1.5 μg of wild type or mutant 211

decarboxylase against 2H1NA over the concentration range of 0.05 to 0.5 mM. The 212

maximum velocity (Vmax) and the Michaelis constant (Km) were determined by Lineweaver-213

Burk double reciprocal plots using GRAPHPAD-PRISM program (version 5.00 for 214

Windows). The optimum temperature of the recombinant protein was determined over the 215

range of 10 to 70°C and the pH profile was determined over the pH range of 4.0-9.5 using the 216

following buffer systems (50 mM): citrate buffer (pH 4-6), sodium phosphate buffer (pH 6-217

8), and glycine-NaOH (pH 8.0-9.5) at the optimal temperature determined above under 218

standard conditions. The effect of temperature on enzyme stability was determined by 219

preincubating the enzyme at different temperatures (10-60°C) for 30 min and measuring the 220

remaining activity under standard conditions. To study the effect of various metal ions and 221

inhibitors on enzyme activity, purified 2H1NA decarboxylase preincubated with respective 222

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

10

metal ions or inhibitors (1 or 5 mM) for 10 min at 4°C, was used as enzyme preparation. 223

However, for metal chelators, viz. EDTA, 1,10-phenanthroline, 2,2′-bipyridyl and 8-224

hydroxyquinoline-5-sulphonic acid (8-HQSA), enzyme was preincubated for 16h. 225

For metal analysis, purified HndA (3 mg) was hydrolyzed by 65% ultrapure concentrated 226

nitric acid (2 ml) (Suprapure, Merck, Darmstadt, Germany) at 110°C for 1 h. The sample was 227

diluted 10-fold by deionized double distilled water, and the metal content was determined by 228

atomic absorption spectrometer (iCE 3000 Series, Thermo Fischer Scientific). 229

Homology modeling and docking analyses. A three-dimensional model of HndA was 230

constructed using 2-amino-3-carboxymuconate-6-semialdehyde decarboxylase (ACMSD) 231

from Pseudomonas fluorescens (PDB ID: 2HBV) (33) as template employing Modeller 9v7 232

(34). The models were checked using Prochek, Verify3D and VADAR (35-37). The NCBI 233

PubChem database (http://pubchem.ncbi.nlm. nih.gov/) was used to obtain co-ordinates of the 234

ligand 2H1NA. Preparation of protein and substrate files (pdbqt files) was performed using 235

AutoDockTools-1.5.6 using default parameters (38). The grid box with dimensions 236

50 × 50 × 50 grid points was generated using AutoGrid4 program keeping the metal ion 237

coordinates (43.43 × -0.323 × 16.82) at the centre. AutoDock4 was used to perform docking 238

using genetic algorithm. Docked poses were analyzed using AutoDockTools-1.5.6 to get the 239

best binding pose of 2H1NA with the lowest binding energy. The binding residues were 240

identified and the schematic diagram of protein-ligand interaction was generated using 241

LigPlot+ suite (version 1.4.5) (39). 242

Nucleotide sequence accession numbers. The nucleotide sequence reported in this paper has 243

been deposited in the DDBJ/EMBL/GenBank database under the accession number 244

KU254672. 245

246

247

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

11

RESULTS 248

Partial purification and gene identification of 2H1NA decarboxylase. 2H1NA 249

decarboxylase activity was previously reported to be strictly inducible in the presence of 250

2H1NA and a differentially expressing ~32 kDa protein band was observed only in the cell-251

free extract of 2H1NA-grown cells when compared to that of 2–naphthol-grown cells (27). 252

For detailed characterization, 2H1NA decarboxylase was partially purified from crude cell-253

free extract of strain BC1 grown on 2H1NA using differential protein precipitation steps and 254

hydrophobic interaction chromatography (Table 1). The purified enzyme preparation was 255

found to be stable during the purification steps, carried out under aerobic conditions 256

indicating oxygen insensitive nature of the decarboxylase. The decarboxylase-active fractions 257

from Phenyl Sepharose column represented 19-fold purification (specific activity 3.8 U mg-1) 258

with a yield of 16.8% and showed the presence of a ~32 kDa band in SDS-PAGE supporting 259

our earlier observation (see Fig. S1 in Supplemental material). 260

To confirm its identity, the ~32 kDa protein was subjected to MALDI-TOF MS/MS 261

analysis where the generated peptide fragments showed strong sequence similarity with the 262

uncharacterized amidohydrolase-2 proteins of various Burkholderia sp. in Blastp analyses 263

(see Table S5 in Supplemental material). Subsequently, a 220 bp PCR product was amplified 264

(data not shown) from the genomic DNA of strain BC1 using the primers, HNDA_F and 265

HNDA_R, which on sequence analysis confirmed the results as stated above. 266

Cloning and sequencing of the 2H1NA decarboxylase gene. Screening of a genomic 267

fosmid library of strain BC1 by PCR led to the identification of a subclone, which upon IPTG 268

induction displayed 2H1NA decarboxylase activity, determined in the cell-free enzyme 269

preparation. Complete sequence analysis of the subclone harbouring a 4.2 kb EcoRI fragment 270

revealed the presence of the amidohydrolase gene designated as hndA for hydroxynaphthoate 271

decarboxylase. The decarboxylase, hndA, consisting of 951 nucleotides, encoded a 272

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

12

polypeptide of 316 amino acids with a theoretical molecular weight of 34 kDa and pI of 5.59. 273

Moreover, the CDD (Conserved Domain Database) and COG (Clusters of Orthologous 274

Groups) analyses of HndA placed it in the amidohydrolase superfamily of the TIM-barrel 275

fold protein (COG2159), which includes several non-oxidative decarboxylases, including 5-276

carboxyvanillate decarboxylase (5-CVD), 2,3-dihydroxybenzoate decarboxylase and γ-277

resorcylate decarboxylase (γ-RSD) (4, 14, 15). HndA showed 71-97% identity with the 278

biochemically uncharacterized metal-dependent hydrolase proteins of several Burkholderia 279

sp. belonging to the amidohydrolase superfamily, listed in the NCBI database. However, 280

among the biochemically well-characterized non-oxidative decarboxylases, HndA showed a 281

modest identity of 27 and 24% with the γ-resorcylate decarboxylase (γ-RSD) of Rhizobium 282

sp. MTP-10005 (15) and 2-amino-3-carboxymuconate-6-semialdehyde decarboxylase 283

(ACMSD) of Pseudomonas fluorescence (10), respectively. Other genes in the 4.2 kb gene 284

cluster includes orf1, orf2 and dbpA where the gene products showed 99-100% identity with 285

a non-characterized phenol degradation protein, a LysR-type regulator and an ATP-286

dependent RNA helicase protein of Burkholderia multivorans respectively. The genetic 287

assembly of the 4.2 kb cluster is shown in Fig. 1A. 288

Phylogenetic analysis of HndA. A phylogenetic tree (Fig. 1B) constructed using multiple 289

sequence alignment of various proteins belonging to the amidohydrolase-1 and 290

amidohydrolase-2 family positioned HndA within the amidohydrolase-2 family. With the 291

exception of 4-oxalomesaconate hydratase (OMAH) from Sphingomonas paucimobilis SYK-292

6 (40), the other representative members belonging to this family are non-oxidative 293

decarboxylases, viz. isoorotate decarboxylase (IDCase) from Neurospora crassa (12) and 5-294

carboxyvanillate decarboxylase (5-CVD) from Sphingomonas paucimobilis SYK-6 (4) apart 295

from γ-RSD from Rhizobium sp. MTP-10005 (15) and ACMSD from Pseudomonas 296

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

13

fluorescence (10). In addition, a number of uncharacterized metal-dependent hydrolases from 297

Burkholderia also appeared to belong to this enzyme family (Fig. 1B). 298

Despite an overall low sequence homology among the biochemically characterized 299

members of amidohydrolase-2 family proteins, the sequence alignment did display strong 300

residue conservation pattern for amino acids (His10, His12, His156 and Asp269 in HndA) 301

that are responsible for the binding of the metal cofactor, crucial for enzyme catalysis (Fig. 302

1C). Apart from above, the alignment showed another conserved histidine residue (His204 in 303

HndA) which was earlier reported to play a crucial role in enzyme catalysis for both ACMSD 304

and γ-RSD (33, 41). 305

Overexpression and purification of recombinant of HndA. The recombinant 2H1NA 306

decarboxylase was successfully overexpressed in E. coli BL21(DE3) under 0.5 mM IPTG 307

concentration and was purified by Ni2+-NTA chromatography (Fig. 2A). The purified 308

recombinant enzyme migrated as a single band in the SDS-PAGE gel with an apparent 309

subunit molecular weight of ~38 kDa. While the molecular weight of the native recombinant 310

enzyme on gel filtration was found to be 38.1±0.5 kDa, suggesting the monomeric nature of 311

the enzyme. Purified recombinant HndA catalyzed the decarboxylation of 2H1NA to 2-312

naphthol as revealed by both spectral and HPLC analyses (Fig. 2B, 2C) with a specific 313

activity of 9.0 U/mg of protein. Fig. 2D shows the time-dependent transformation of 2H1NA 314

to 2-naphthol by purified HndA over a period of 10 minutes. 315

Biochemical properties of recombinant HndA. It was observed that HndA did not favour 316

carboxylation of 2-naphthol under the conditions tested and thus it appears that the enzyme 317

catalyzes an irreversible reaction (decarboxylation). Again, HndA showed strict substrate 318

specificity towards 2H1NA since its other structural isomers, 1H2NA and 3H2NA, could not 319

be transformed. Also, it failed to decarboxylate mono- and dihydroxybenzoic acids, viz. 2-320

hydroxybenzoic acid (salicylic acid), 3-hydroxybenzoic acid, 4-hydroxybenzoic acid, 2,3-321

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

14

dihydroxybenzoic acid, 2,4-dihydroxybenzoic acid, 2,5-dihydroxybenzoic acid (gentisic acid) 322

and 2,6-dihydroxybenzoic acid (γ-resorcylic acid). Similarly, HndA failed to transform 323

phthalic acid, 1-naphthoic acid and 2-naphthoic acid. 324

For 2H1NA decarboxylase, the values of Km and Vmax were determined to be 0.17 mM and 325

0.02 μmol/min, respectively. The kcat/Km value for 2H1NA was 47.05 mM-1 s-1. The optimum 326

pH and temperature of the protein was found to be 7.5 and 35°C respectively (see Fig. S2A 327

and S2B in Supplemental material). The enzyme was found to be stable up to 45°C and 328

retained 58% of initial activity when incubated at 50°C for 30 min. However, the enzyme 329

completely lost its activity when incubated above 60°C (see Fig. S2C in Supplemental 330

material). 331

The effect of various metal ions as well as inhibitors on enzyme activity is shown in 332

supplementary Table S6. No significant change in enzyme activity was observed with 333

majority of the metal ions, inhibitors and metal ion chelators, individually incubated for 10 334

min. However, activity of the enzyme was inhibited by AgNO3 and HgCl2, as suggested for 335

γ-RSD (15, 42). Nevertheless, a modest inhibition in HndA activity was observed when the 336

enzyme was incubated for 16 h individually with metal chelators including 8-hydroxy-337

quinoline-5-sulfonic acid, a zinc metal specific inhibitor. This observation suggests the 338

possible presence of a deeply embedded metal ion, inadequately accessible by metal 339

chelators. Diethylpyrocarbonate, a histidine residue modifier also showed a reasonable 340

decrease in enzymatic activity suggesting the presence of active site histidine residues in 341

HndA as described earlier (28). To identify the metal centre in the active site of HndA, 342

atomic absorption spectroscopy analysis was performed which revealed the presence of zinc 343

at 0.95±0.1 mol per mol of protein. This result corroborated well with other non-oxidative 344

decarboxylases belonging to the amidohydrolase superfamily that possess zinc as the 345

transition metal centre (33, 41). 346

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

15

Structural modelling and functional analysis of HndA mutants. A three-dimentional 347

(3D) structural model of HndA showed the presence of a (β/α)8 barrel fold with eight parallel 348

β strands flanked by eight α helices on the outer face. The structural proximity of the 349

conserved His10, His12, His156 and Asp269 residues in the 3D model of HndA was fully 350

compatible with their putative role in forming a metal ion binding motif (Fig. 3A). In order to 351

confirm the functional roles of these amino acid residues, site-directed mutants HndAH10Q, 352

HndAH156Q and HndAD269V were constructed, expressed as soluble proteins (see Fig. S3 in 353

Supplemental material) and their specific activity against 2H1NA were determined. 354

HndAH10Q and HndAD269V showed no activity against 2H1NA whereas HndAH156Q retained 355

only 7.23% (0.65 U/mg) of the wild type decarboxylase activity. To analyse the role of the 356

conserved histidine residue His204, the mutant protein HndAH204Q was studied which showed 357

a complete loss of enzymatic activity, suggesting a critical role of this residue in HndA-358

mediated catalysis. 359

In order to determine other important amino acid residues responsible for substrate 360

binding, docking analysis was performed using 2H1NA as a ligand. Docking analysis 361

revealed the role of two important amino acid residues, Arg33 and Tyr272, which were found 362

to interact with the carboxyl functional group of 2H1NA by hydrogen bonding (Fig. 3B). In 363

addition to the carboxyl group, Tyr272 was also found to interact with the hydroxyl group of 364

2H1NA (Fig. 3B). To further assess their role in substrate binding, HndAR33L and HndAY272F 365

mutant proteins were generated (see Fig. S3 in Supplemental material) and subsequently their 366

activities were tested against 2H1NA. The mutant enzymes, HndAR33L and HndAY272F, 367

showed specific activities of 4.1 and 2.8 U/mg of protein respectively. Kinetic parameters 368

determined for these mutants showed negligible change in Km for HndAH156Q but showed a 369

clear increase in Km for mutants HndAR33L and HndAY272F, suggesting their role as substrate 370

binding residues. Also, the kcat/Km values for the mutants decreased by nearly six to fifteen-371

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

16

fold with respect to wild type protein indicating that all of these residues are essential for 372

2H1NA decarboxylation reaction (Table 2). 373

374

DISCUSSION 375

Decarboxylation is ubiquitous in nature and is of fundamental biological importance. Among 376

the different classes of decarboxylases, nonoxidative decarboxylases have received specific 377

attention primarily because they catalyze transition metal-dependent decarboxylation without 378

using molecular oxygen as a cosubstrate (1). Microorganisms expressing these enzymes not 379

only play a significant role in biodegradation and/or bioremediation of soil, water and 380

sediment-contaminated with lignin-related compounds and benzene-derivatives of industrial 381

origin but also act as biocatalysts in industrial biotransformation reactions (4, 15, 17). 382

Non-oxidative decarboxylases belonging to structurally distinct protein families differ in 383

their oxygen sensitivity that catalyze either reversible or irreversible reactions (42). In 384

aromatic acid metabolism, presence of a variety of hydroxybenzoic acid non-oxidative 385

decarboxylases has been detected and some have been purified and characterized (2, 5, 14, 386

15, 42). To the best of our knowledge, the oxygen insensitive 2H1NA non-oxidative 387

decarboxylase described in this study is the first bacterial enzyme belonging to the 388

amidohydrolase superfamily that catalyzes an irreversible decarboxylation of a 389

hydroxynaphthoic acid. 390

The amidohydrolase superfamily is comprised of functionally diverse enzymes that 391

catalyze the cleavage of C–N, C–C, C–O, C–Cl, C–S or O–P bond of structurally distinct 392

organic compounds (43-45). Generally, members of amidohydrolase superfamily share a 393

signature for mono- or binuclear metal centre embedded within the triosephosphate isomerase 394

(TIM)-like barrel fold in the catalytic domain (43, 44). Within the amidohydrolase 395

superfamily, members of the amidohydrolase-1 family catalyze hydrolytic reactions while the 396

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

17

amidohydrolase-2 family proteins are primarily involved in non-hydrolytic C-C bond 397

cleavage (44). The gene encoding the 2H1NA decarboxylase (hndA) displayed similarities 398

with the members of the amidohydrolase-2 family. In this family, ACMSD from 399

Pseudomonas fluorescence is the first characterized member involved in 2-nitrobenzoic acid 400

degradation pathway (10). Other members belonging to this family include 5CVD, γ-RSD, 401

salicylic acid decarboxylase, OMAH and Idcase (4, 15, 17, 40, 46). Since none of the 402

enzymes belonging to amidohydrolase-2 family possess hydrolase activity, Liu & Zhang (43) 403

had proposed to rename this family as ACMSD-related protein family. Based on sequence 404

similarity and phylogenetic analysis (Fig. 1), we propose HndA to be a new member of 405

ACMSD-related protein family. 406

Multiple sequence alignment of ACMSD-related protein family members, including 407

HndA, revealed strict conservation pattern for key amino acid residues which act as important 408

metal binding protein ligands (43). For HndA, His10 and His12 constitute the conserved 409

“HxH” metal binding motif whereas H156 and D269 are the other two endogenous metal 410

binding ligands (Fig. 3A). Enzyme inhibition by histidine residue specific inhibitor 411

diethylpyrocarbonate (DEPC) and site directed mutagenesis studies revealed that the 412

conserved histidine residues constitute the active site protein ligands (Table 2). Interestingly, 413

substitution of another conserved histidine residue (His204), not directly involved in metal 414

binding, leads to the complete inactivation of the protein. This result is similar to that 415

observed in ACMSD where the corresponding residue (His228) was suggested to play the 416

role of an acid-base catalyst involved in deprotonation of the metal-bound water facilitating 417

the decarboxylation of ACMS (47). 418

On the other hand, docking analysis revealed that carboxyl and hydroxyl groups of 419

2H1NA are hydrogen bonded with Arg33/Tyr272 and Tyr272 of HndA, respectively. 420

Interestingly, the importance of arginine residue in substrate binding via carboxylate group 421

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

18

has been suggested in ACMSD (33, 48). Similarly, the role of the active site tyrosine residue 422

in the binding of the hydroxyl group of lactic acid was studied in flavocytochromeb2 or L-423

lactate dehydrogenase and it was reported to play a role in converting lactic acid to pyruvic 424

acid (49). The significance of Arg33 and Tyr 272 for efficient substrate binding in HndA was 425

also confirmed by mutational analysis (Table 2). 426

Being a member of metallo-dependent hydrolase superfamily, HndA did not show any 427

enhancement in activity when a set of metal ions were individually supplemented externally. 428

Again, common divalent metal chelators had only mild inhibitory effects on this enzyme (see 429

Table S6 in supplemental material) even after prolonged incubation, suggesting a probable 430

deeply buried metal centre within the protein molecule. Modest inhibition by a zinc metal-431

specific inhibitor, 8-hydroxy-quinoloine-5-sulphonic acid (see Table S6 in supplemental 432

material) that suggested the possible presence of a zinc metal centre within the enzyme was 433

verified by atomic absorption spectroscopy. Additional biophysical investigations on HndA 434

will provide further insights on the catalytic mechanism and structure–function relationships 435

of this unique transition metal-dependent, oxygen-independent 2H1NA decarboxylase. 436

437

ACKNOWLEDGMENTS 438

Authors acknowledge Gautam Basu for editing the manuscript. Financial support for this 439

work was provided by Bose Institute, Kolkata, India. P.P.C. was supported with fellowships 440

from the Council of Scientific and Industrial Research, Government of India while S.B. and 441

A.D were supported with fellowships from Bose Institute. 442

443

REFERENCES 444

1. Li T, Huo L, Pulley C, Liu A. 2012. Decarboxylation mechanisms in biological system. 445

Bioorg Chem 43: 2-14. 446

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

19

2. Huang J, He Z, Wiegel J. 1999. Cloning, characterization, and expression of a novel 447

gene encoding a reversible 4-hydroxybenzoate decarboxylase from Clostridium 448

hydroxybenzoicum. J Bacteriol 181: 5119-5122. 449

3. Chow KT, Pop MK, Davies J. 1999. Characterization of a vanillic acid non-oxidative 450

decarboxylation gene cluster from Streptomyces sp. D7. Microbiology 145: 2393-2403. 451

4. Peng X, Masai E, Kitayama H, Harada K, Katayama Y, Fukuda M. 2002. 452

Characterization of the 5-carboxyvanillate decarboxylase gene and its role in lignin-453

related biphenyl catabolism in Sphingomonas paucimobilis SYK-6. Appl Environ 454

Microbiol 68: 4407-4415. 455

5. He Z, Wiegel J. 1996. Purification and characterization of an oxygen-sensitive, 456

reversible 3,4-dihydroxybenzoate decarboxylase from Clostridium hydroxybenzoicum. J 457

Bacteriol 178: 3539-3543. 458

6. Gu W, Li X, Huang J, Duan Y, Meng Z, Zhang KQ, Yang J. 2011. Cloning, 459

sequencing, and overexpression in Escherichia coli of the Enterobacter sp. Px6-4 gene 460

for ferulic acid decarboxylase. Appl Microbiol Biotechnol 89: 1797-1805. 461

7. Cavin JF, Barthelmebs L, Divies C. 1997. Molecular characterization of an inducible 462

p-coumaric acid decarboxylase from Lactobacillus plantarum: gene cloning, 463

transcriptional analysis, overexpression in Escherichia coli, purification, and 464

characterization. Appl Environ Microbiol 63: 1939-1944. 465

8. Pujar BG, Ribbons DW. 1985. Phthalate metabolism in Pseudomonas fluorescens 466

PHK: purification and properties of 4,5-dihydroxyphthalate decarboxylase. Appl Environ 467

Microbiol 49: 374-376. 468

9. Muraki T, Taki M, Hasegawa Y, Iwaki H, Lau PCK. 2003. Prokaryotic homologs of 469

the eukaryotic 3-hydroxyanthranilate 3,4-dioxygenase and 2-amino-3-carboxymuconate-470

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

20

6-semialdehyde decarboxylase in the 2-nitrobenzoate degradation pathway of 471

Pseudomonas fluorescens strain KU-7. Appl Environ Microbiol 69: 1564-1572. 472

10. Li T, Iwaki H, Fu R, Hasegawa Y, Zhang H, Liu A. 2006. α-Amino-β-473

carboxymuconic-ε-semialdehyde decarboxylase (ACMSD) is a new member of the 474

amidohydrolase superfamily. Biochemistry 45: 6628-6634. 475

11. Colabroy KL, Begley TP. 2005. Tryptophan catabolism: identification and 476

characterization of a new degradative pathway. J Bacteriol. 187: 7866-7869. 477

12. Smiley JA, Kundracik M, Landfried DA, Barnes V R Sr., Axhemi AA. 2005. Genes 478

of the thymidine salvage pathway: thymine-7-hydroxylase from a Rhodotorula glutinis 479

cDNA library and isoorotate decarboxylase from Neurospora crassa. Biochim Biophys 480

Acta 1723: 256-264. 481

13. Yoshida T, Fujita K, Nagasawa T. 2002. Novel reversible indole-3-carboxylate 482

decarboxylases catalyzing non-oxidative decarboxylation. Biosci Biotechnol Biochem. 483

66: 2388-2394. 484

14. Santha R, Savithri HS, Rao NA, Vaidyanathan CS. 1995. 2,3-Dihydroxybenzoic acid 485

decarboxylase from Aspergillus niger. A novel decarboxylase. Eur J Biochem 230: 104-486

110. 487

15. Yoshida M, Fukuhara N, Oikawa T. 2004. Thermophilic, reversible γ-resorcylate 488

decarboxylase from Rhizobium sp. strain MTP-10005: Purification, molecular 489

characterization and expression. J Bacteriol 186: 6855-6863. 490

16. Lupa B, Lyon D, Shaw LN, Sieprawska-Lupa M, Wiegel J. 2008. Properties of the 491

reversible nonoxidative vanillate/4-hydroxybenzoate decarboxylase from Bacillus 492

subtilis. Can J Microbiol 54: 75-81. 493

17. Kirimura K, Gunji H, Wakayama R, Hattori T, Ishii Y. 2010. Enzymatic Kolbe-494

Schmitt reaction to from salicylic acid from phenol: enzymatic characterization and gene 495

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

21

identification of a novel enzyme, Trichosporan moniliforme salicylic acid decarboxylase. 496

Biochem Biophys Res Commun 394: 279-284. 497

18. Iwabuchi T, Harayama S. 1998. Biochemical and molecular characterization of 1-498

hydroxy-2-naphthoate dioxygenase from Nocardioides sp. KP7. J Biol Chem 273: 8332-499

8336. 500

19. Vila J, Lopez Z, Sabate J, Minguillon C, Solanas AM, Grifoll M. 2001. Identification 501

of a novel metabolite in the degradation of pyrene by Mycobacterium sp. Strain AP1: 502

Actions of the isolate on two- and three-ring polycyclic aromatic hydrocarbons. Appl 503

Environ Microbiol 67: 5497-5505. 504

20. Mallick S, Chatterjee S, Dutta TK. 2007. A novel degradation pathway in the 505

assimilation of phenanthrene by Staphylococcus sp. strain PN/Y via meta-cleavage of 2-506

hydroxy-1-naphthoic acid: formation of trans-2,3-dioxo-5-(2′-hydroxyphenyl)-pent-4-507

enoic acid. Microbiology 153: 2104-2115. 508

21. Kanaly RA, Harayama S. 2000. Biodegradation of high-molecular-weight polycyclic 509

aromatic hydrocarbons by bacteria. J Bacteriol 182: 2059–2067. 510

22. Kanaly RA, Harayama S. 2010. Advances in the field of high-molecular-weight 511

polycyclic aromatic hydrocarbon biodegradation by bacteria. Microb Biotechnol 3: 136-512

164. 513

23. Pinyakong O, Habe H, Supaka N, Pinpanichkarn P, Juntongjin K, Yoshida T, 514

Furihata K, Nojiri H, Yamane H, Omori T. 2000. Identification of novel metabolites 515

in the degradation of phenanthrene by Sphingomonas sp. strain P2. FEMS Microbiol 516

Lett 191: 115-121. 517

24. Tao XQ, Lu GN, Dang Z, Yang C,Yi XY. 2007. A phenanthrene-degrading strain 518

Sphingomonas sp. GY2B isolated from contaminated soils. Process Biochem 42: 401-519

408. 520

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

22

25. Prabhu Y, Phale PS. 2003. Biodegradation of phenanthrene by Pseudomonas sp. strain 521

PP2: novel metabolic pathway, role of biosurfactant and cell surface 522

hydrophobicity in hydrocarbon assimilation. Appl Microbiol Biotechnol 61: 342-51. 523

26. Samanta SK, Chakraborti AK, Jain RK. 1999. Degradation of phenanthrene by 524

different bacteria: evidence for novel transformation sequences involving the formation 525

of 1–naphthol. Appl Microbiol Biotechnol 53: 98-107. 526

27. Chowdhury PP, Sarkar J, Basu S, Dutta TK. 2014. Metabolism of 2-hydroxy-1-527

naphthoic acid and naphthalene via gentisic acid by distinctly different sets of enzymes 528

in Burkholderia sp. strain BC1. Microbiology 160: 892-902. 529

28. Bradford MM. 1976. A rapid and sensitive method for the quantitation of protein 530

utilizing the principle of protein-dye binding. Anal Biochem 72: 248-254. 531

29. Shevchenko A, Tomas, H, Havlis J, Olsen JV, Mann M. 2006. In-gel digestion for 532

mass spectrometric characterization of proteins and proteomes. Nat Protoc 1: 2856-2860. 533

30. Altschul SF, Gish W, Miller W, Myers EW, Lipman DJ. 1990. Basic local alignment 534

search tool. J Mol Biol 215: 403-410. 535

31. Thompson JD, Gilson TJ, Plewniak F, Jeanmougin F. Higgins DG. 1997. The 536

ClustalX windows interface: flexible strategies for multiple sequence alignment aided by 537

quality analysis tools. Nucleic Acids Res 24: 4876‐4882. 538

32. Tamura K, Peterson D, Peterson N, Stecher G, Nei M, Kumar S. 2011. MEGA 5: 539

Molecular evolutionary genetics analyses using maximum likelihood, evolutionary 540

distance and maximum parsimony methods. Mol Biol Evol 28: 2731-2739. 541

33. Martynowski D, Eyobo Y, Li T, Kun Yang K, Liu A, Zhang H. 2006. Crystal 542

structure of α-amino-β-carboxymuconate-ε-semialdehyde decarboxylase: Insight into the 543

active site and catalytic mechanism of a novel decarboxylation reaction. Biochemistry 544

45: 10412-10421. 545

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

23

34. Sali A, Blundell TL. 1993. Comparative protein modelling by satisfaction of spatial 546

restraints. J Mol Biol 234: 779–815. 547

35. Luthy R, Bowie JU, Eisenberg D. 1992. Assessment of protein models with three-548

dimensional profiles. Nature 356: 83–85. 549

36. Laskowski RA, MacArthur MW, Moss DS, Thornton JM. 1993. Procheck: a 550

program to check the stereochemical quality of protein structures. J Appl Crystallogr 26: 551

283-291. 552

37. Willard L, Ranjan A, Zhang H, Monzavi H, Boyko RF, Sykes BD, Wishart DS. 553

2003. VADAR: a web server for quantitative evaluation of protein structure quality. 554

Nucleic Acids Res 31: 3316-3319. 555

38. Morris G M, Huey R, Lindstrom W, Sanner MF, Belew RK, Goodsell DS, Olson 556

AJ. 2009. Autodock4 and AutoDockTools4: automated docking with selective receptor 557

flexiblity. J Comput Chem 30: 2785-2791. 558

39. Laskowski R A, Swindells M B. 2011. LigPlot+: multiple ligand-protein interaction 559

diagrams for drug discovery. J Chem Inf Model. 51: 2778–2786. 560

40. Hara H, Masai E, Katayama Y, Fukuda M. 2000. The 4-oxalomesaconate hydratase 561

gene, involved in the protocatechuate 4,5-cleavage pathway, is essential to vanillate and 562

syringate degradation in Sphingomonas paucimobilis SYK-6. J Bacteriol 182: 6950-563

6957. 564

41. Goto M, Hayashi H, Miyahara I, Hirotsu K, Yoshida M, Oikawa T. 2006. Crystal 565

structures of nonoxidative zinc-dependent 2,6-dihydroxybenzoate (γ-resorcylate) 566

decarboxylase from Rhizobium sp. strain MTP-10005. J Biol Chem 281: 34365-34373. 567

42. Ishii Y, Narimatsu Y, Iwasaki Y, Arai N, Kino K, Kirimura K. 2004. Reversible and 568

nonoxidative γ-resorcylic acid decarboxylase: Characterization and gene cloning of a 569

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

24

novel enzyme catalyzing carboxylation of resorcinol, 1,3-dihydroxybenzene, from 570

Rhizobium radiobacter. Biochem Biophys Res Commun 324: 611-620. 571

43. Liu A, Zhang H. 2006. Transition metal-catalyzed nonoxidative decarboxylation 572

Reactions. Biochemistry 45: 10407-10411. 573

44. Seibert CM, Raushal FM. 2005. Structural and catalytic diversity within the 574

amidohydrolase superfamily. Biochemistry 44: 6384-6391. 575

45. Kim GJ, Lee DE, Kim HS. 2000. Functional expression and characterization of the two 576

cyclic amidohydrolase enzymes, allantoinase and a novel phenyl hydantoinase, from 577

Escherichia coli. J Bacteriol 182: 7021-7028. 578

46. Xu S, Li W, Zhu J, Wang R, Li Z, Xu GL, Ding J. 2013. Crystal structures of 579

isoorotate decarboxylases reveal a novel catalytic mechanism of 5-carboxyl-uracil 580

decarboxylation and shed light on the search for DNA decarboxylase. Cell Res 23: 1296-581

1309. 582

47. Huo L, Fielding AJ, Chen Y, Li T, Iwaki H, Hosler JP, Chen L, Hasigawa Y, Que L 583

Jr, Liu A. 2012. Evidence for a dual role of an active site histidine in α-amino-β-584

carboxymuconate-ε-semialdehyde decarboxylase. Biochemistry 51: 5811-5821. 585

48. Huo L, Davis I, Chen L, Liu A. 2013. The power of two: arginine 51 and arginine 239* 586

from a neighboring subunit are essential for catalysis in α-amino-β-carboxymuconate-ε-587

semialdehyde decarboxylase. J Biol Chem 288: 30862-30871. 588

49. Gondry M, Dubois J, Terrier M, Lederer F. 2001. The catalytic role of tyrosine 254 in 589

flavocytochrome b2 (L-lactate dehydrogenase from baker's yeast). Comparison between 590

the Y254F and Y254L mutant proteins. Eur J Biochem 268: 4918-4927. 591

592

FIGURE LEGEND 593

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

25

FIG 1 (A) Gene organization of 4.2 kb EcoRI fragment showing gene designations. orf1, 594

partial phenol degradation protein; hndA, 2H1NA decarboxylase; orf2, LysR-type like 595

regulator; dbpA, ATP-dependent RNA helicase. (B) Phylogenetic tree based on protein 596

sequences from amidohydrolase-1 and amidohydrolase-2 family of proteins. Numbers at 597

nodes indicate levels of bootstrap support based on neighbour joining analysis of 100 598

resampled data sets. Bootstrap values below 50% are not shown. Bar represents 0.1 599

substitutions per nucleotide position. GenBank or PDB accession numbers are indicated 600

within parentheses. Amidohydrolase-3 from Burkholderia sp. lig30 (KDB07616) was used as 601

an outgroup. (C) Multiple sequence alignment of protein sequences of the representative 602

members of amidohydrolase-2 protein family. Metal coordinating residues are shaded. 603

604

FIG 2 (A) SDS PAGE analysis of overexpressed recombinant HndA protein. Lane 1, crude 605

extract of E. coli BL21(DE3) carrying empty pET28a vector; lane 2, crude extracts of 606

induced E. coli BL21(DE3) carrying pET28a:HndA; lane 3, purified recombinant HndA 607

protein and lane M, molecular weight marker (Puregene, Genetix, India). (B) Spectral 608

changes during transformation of 2HINA by purified recombinant HndA protein. The sample 609

and reference cuvettes contained 50 mM potassium phosphate buffer (pH 7.0) in 1-ml 610

volume. The sample cuvette also contained 220 nmol 2H1NA. Spectra were recorded every 611

one minute after the addition of 10 µg protein to both cuvettes. The up and down arrows 612

indicate increasing and decreasing absorbance, respectively. (C) HPLC chromatogram 613

showing transformation of 2H1NA to 2-naphthol by purified HndA in a reaction mixture 614

(final volume, 1 ml) containing 0.5 mM 2H1NA and 5 µg protein in buffer A incubated for 615

10 min at 35°C. UV-visible spectra of 2H1NA and 2-naphthol are shown in insets. (D) Time-616

dependent transformation of 2H1NA to 2-naphthol by purified HndA. Concentration of 617

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

26

2H1NA (○) and 2-naphthol (●) were determined by HPLC from the reaction mixtures (as 618

described in panel C) during enzymatic transformation over 10 minutes. 619

620

FIG 3 (A) Schematic representation of the structural model of HndA showing the enzyme 621

active site. Inset shows the metal coordinating residues His10, His12, His156 and Asp269. 622

(B) Surface topology of HndA showing the binding of 2H1NA within the catalytic pocket via 623

electrostatic interaction with active site residues Tyr272 and Arg33 based on docking 624

analysis. 625

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

TABLE 1 Purification summary of 2H1NA decarboxylase from Burkholderia sp. BC1

Step

Total

protein

(mg)

Total

activity

(U)

Specific

activity

(U/mg)

Purification

(fold)

Yield

(%)

Cell-free extract 180 36.3 0.2 1 100

Ammonium sulphate

fractionation (30-50%

saturation)

32

25.9

0.8

4

71.3

Phenyl-sepharose 6 (Fast

Flow) 1.6 6.1 3.8 19 16.8

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from

1

TABLE 2 Kinetic constants of wild type and mutant 2H1NA decarboxylases

Enzyme Km

(mM) Vmax

(μmol min-1) kcat (s-1)

kcat/Km (mM-1 s-1)

HnD 0.17 0.02 7.99 47.27

HnDH10Q ND ND ND ND

HnDH156Q 0.17 0.001 0.56 3.31

HnDD269V ND ND ND ND

HnDH204Q ND ND ND ND

HnDR33L 0.78 0.01 6.33 8.06

HnDY272F 0.31 0.007 3.02 9.71

ND, Not determined (as product could not be detected).

on April 16, 2018 by guest

http://jb.asm.org/

Dow

nloaded from