38
1 X-ray structure of the amidase domain of AtzF, the allophanate hydrolase from the 1 cyanuric acid-mineralizing multienzyme complex 2 3 Sahil Balotra, 1,2 Janet Newman, 3 Nathan P. Cowieson 4 , Nigel G. French, 1 Peter M. Campbell, 1 4 Lyndall J. Briggs, 1 Andrew C. Warden, 1 Christopher J. Easton, 2 Thomas S. Peat, 3 Colin Scott 1 5 6 1 CSIRO Land and Water Flagship, Black Mountain, Canberra, Australia, 2601 7 2 Research School of Chemistry, Australian National University, Canberra, 2601 8 3 CSIRO Biomedical Manufacturing Program, Parkville, Australia, 3052 9 4 Australian Synchrotron, 800 Blackburn Rd, Clayton VIC 3168, Australia 10 11 Correspondence should be addressed to [email protected] 12 13 Key words: substrate channeling; Ser-cisSer-Lys; atrazine; protein-protein interaction 14 AEM Accepts, published online ahead of print on 31 October 2014 Appl. Environ. Microbiol. doi:10.1128/AEM.02783-14 Copyright © 2014, American Society for Microbiology. All Rights Reserved. on October 17, 2020 by guest http://aem.asm.org/ Downloaded from

on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

  • Upload
    others

  • View
    1

  • Download
    0

Embed Size (px)

Citation preview

Page 1: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

1

X-ray structure of the amidase domain of AtzF, the allophanate hydrolase from the 1

cyanuric acid-mineralizing multienzyme complex 2

3

Sahil Balotra,1,2 Janet Newman,3 Nathan P. Cowieson4, Nigel G. French,1 Peter M. Campbell,1 4

Lyndall J. Briggs,1 Andrew C. Warden,1 Christopher J. Easton,2 Thomas S. Peat,3 Colin Scott1 5

6

1CSIRO Land and Water Flagship, Black Mountain, Canberra, Australia, 2601 7

2Research School of Chemistry, Australian National University, Canberra, 2601 8

3CSIRO Biomedical Manufacturing Program, Parkville, Australia, 3052 9

4Australian Synchrotron, 800 Blackburn Rd, Clayton VIC 3168, Australia 10

11

Correspondence should be addressed to [email protected] 12

13

Key words: substrate channeling; Ser-cisSer-Lys; atrazine; protein-protein interaction 14

AEM Accepts, published online ahead of print on 31 October 2014Appl. Environ. Microbiol. doi:10.1128/AEM.02783-14Copyright © 2014, American Society for Microbiology. All Rights Reserved.

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 2: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

2

Abstract 15

The allophanate hydrolase from Pseudomonas sp. strain ADP, AtzF, provides the final 16

hydrolytic step for the mineralization of s-triazines such as atrazine and cyanuric acid. 17

Indeed, the action of AtzF provides metabolic access to two of the three nitrogens in each 18

triazine ring. The X-ray structure of the N-terminal amidase domain of AtzF reveals that it is 19

highly homologous to allophanate hydrolases involved in a different catabolic process in 20

other organisms (i.e. the mineralization of urea). The smaller C-terminal domain does not 21

appear to have a physiologically-relevant catalytic function, as reported for the allophanate 22

hydrolase of Kluyveromyces lactis, when purified enzyme was tested in vitro. However, the 23

C-terminal domain does have a function in coordinating the quaternary structure of AtzF. 24

Interesting, we also show that AtzF forms a large, ca. 660 kDa, multi-enzyme complex with 25

AtzD and AtzE that is capable of mineralizing cyanuric acid. The function of this complex 26

may be to channel substrates from one active site to the next, effectively protecting 27

unstable metabolites, such as allophanate, from solvent-mediated decarboxylation to a 28

‘dead-end’ metabolic product. 29

30

31

32

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 3: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

3

Introduction 33

Atrazine (1-chloro-3-ethylamino-5-isopropylamino-2,4,6-triazine; Fig. 1) is one of the 34

most heavily applied herbicides in the world and is registered for use in North and South 35

America, Australia, Africa, Asia and the Middle East. Atrazine is environmentally persistent 36

(half-life from 4-57 weeks, depending on location) and mobile, leading to the detection of 37

atrazine in surface water, ground water and aquifers (1-3). Atrazine has been detected in 38

the environment at concentrations of up to 4.6 μM in several countries (2, 3). It has been 39

suggested that atrazine may be a carcinogen and an endocrine disrupter at such 40

concentrations (4-6). 41

Since atrazine was introduced in the 1950s, bacteria have evolved highly efficient 42

catabolic pathways that allow the use of atrazine as a sole nitrogen and carbon source (7-43

10). These pathways have provided valuable insights into the evolutionary processes that 44

drive the establishment of new enzyme function and new catabolic pathways (11-15). In 45

addition, these pathways and cognate enzymes provide a potential biotechnological 46

solution to atrazine contamination (i.e. bioremediation; (16-19)). 47

The most intensively studied atrazine-catabolism pathway was discovered in 48

Pseudomonas sp. strain ADP in the mid-1990s, and is comprised of six hydrolases: Atrazine 49

chlorohydrolase (AtzA; EC 3.8.1.8; (20, 21)), N-ethylaminohydrolase (AtzB; EC 3.5.99.3; (22, 50

23)), N-isopropylammelide isopropylaminohydrolase (AtzC; EC 3.5.99.4; (24, 25)), cyanuric 51

acid amidohydrolase (AtzD; EC 3.5.2.15; (15, 26, 27)), biuret amidohydrolase (AtzE; EC 52

3.5.1.84; (28)) and allophanate hydrolase (AtzF; EC 3.5.1.54; (29-31)). These hydrolases 53

sequentially dechlorinate (AtzA) and remove the two N-alkyl side groups (AtzB and AtzC) to 54

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 4: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

4

produce cyanuric acid, which is then further hydrolyzed to biuret, allophanate and ammonia 55

via AtzD, AtzE and AtzF, respectively (Fig. 1). 56

In other bacterial genera, such as Arthrobacter and Nocoidiodes, the function of AtzA 57

is conducted by TrzN (32-36). Although TrzN is physiologically analogous to AtzA, it appears 58

to have evolved independently as the two enzymes share low sequence identity and have 59

significantly different reaction mechanisms and substrate ranges. 60

Recently, there have been concerted efforts to understand the structural biology of 61

the hydrolases of the atrazine catabolic pathway, with the structure of TrzN obtained in 62

2010 (33) and the structure of AtzD (15), which possesses a previously unreported 63

‘Toblerone’ protein fold, reported in 2013. A structure for AtzC has also been deposited in 64

the PDB (2QT3). Structures for AtzB, AtzE and AtzF have yet to be reported, although 65

structures for the ureolytic allophanate hydrolases (AH) from Kluyveromyces lactis (AHKl) 66

and Granulibacter bethesdensis (AHGb) were reported in 2013 (37, 38). Unlike the atrazine-67

degrading AHs, the ureolytic enzymes are found as multi-domain enzymes along with a 68

biotin- and ATP-dependent urea carboxylase, which is required to generate allophanate 69

from urea (37, 39-42). The AH component of these complexes have considerable sequence 70

conservation with each other and the triazine-related AHs (Fig. 2). 71

AH is a member of the amidase signature family, characterized by a ~130 amino acid 72

long region that is rich in serine and glycine residues and contains a characteristic Ser-cisSer-73

Lys catalytic triad (43, 44). Uniquely among this family of amidases, the AHs possess a 74

conserved ca. 15 kDa extension at the C-terminus of the amidase domain, of uncertain 75

function, that is present in both eukaryotic and prokaryotic AHs (38). In a report from Fan et 76

al., the authors proposed that the C-terminal domain of AHKl is involved in the hydrolysis of 77

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 5: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

5

N-carboxycarbamate, the unstable product of AH-mediated allophanate deamination (Fig. 1; 78

(37)). Moreover, in silico substrate docking suggested that a C-terminal histidine residue 79

(His492) plays an important catalytic role (37). 80

Herein, we present the X-ray structure of the N-terminal amidase domain of AtzF 81

from Pseudomonas sp. strain ADP and compare it with the recently determined AH 82

structures. Our in vitro biochemical data were unable to support a physiologically-relevant 83

catalytic function for the C-terminal domain of AtzF. However, we present evidence for a 84

functional role for the ~15 kDa C-terminal domain in coordinating the quaternary structure 85

of the protein; AtzF forms a large multienzyme complex with AtzD and AtzE that mineralizes 86

cyanuric acid. The function of this complex may be to protect unstable metabolites, such as 87

allophanate, from solvent-mediated decarboxylation to a ‘dead-end’ metabolic product. 88

89

Methods and Materials 90

DNA manipulation - The cloning of atzF and the truncated gene encoding the amidase (N-91

terminal) domain of AtzF (atzF467) is described elsewhere (45). . A mutant of atzF that 92

encoded the AtzF H488A variant was produced by the over-lapping PCR method of Ho et al. 93

(46), using atzF as template DNA and atzF H488A fwd (5’ 94

ACCAGCCCTTGAATGCTCAGCTCACGGAG 3’) and atzF H488A rev (5’ 95

TCTCCGTGAGCTGAGCATTCAAGGGCTGG 3’) primers. PCRs were conducted under the 96

following conditions: 30 seconds denaturation (98°C), 30 seconds annealing (55°C) followed 97

by 2 minute extension (72°C) and used Phusion DNA polymerase (NEB, Massachusetts, USA). 98

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 6: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

6

The atzF H488A amplicon was digested with NdeI and BamHI (NEB, Massachusetts, 99

United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 100

was used to transform electrocompetent E. coli BL21 λ(DE3) (Invitrogen, California) 101

following the manufacturer’s instructions. 102

103

Protein expression and purification - Expression, post-expression harvest and lysis of cells for 104

the AtzF H488A variant was carried out under the same conditions used for AtzF and AtzF467 105

expression reported elsewhere (45). Purification of AtzF H488A from soluble cell-free 106

extract was carried out by metal ion affinity chromatography using a Ni–NTA Superflow 107

cartridge (Qiagen, Maryland). After loading the soluble fraction onto the column, a step 108

gradient of 12.5 mM imidazole was applied for six column volumes, followed by a seven 109

column volume wash with 50 mM imidazole after which protein was eluted in seven column 110

volumes of 250 mM imidazole. 111

112

Crystallization and structure solution - Crystallization and data collection were done as 113

previously described (45). Briefly, crystals grew from a reservoir containing 11–14% (w/v) 114

PEG 3350, 2% Tacsimate pH 5 at 293 K, and were used to collect X-ray data at the MX-2 115

beamline of the Australian Synchrotron using a wavelength of 0.9529 Å (13,011 eV). The 116

structure was solved using molecular replacement (Phaser; (47)) with PDB entry 2DQN (48) 117

and a clear solution was found with six protomers in the asymmetric unit. The spacegroup 118

was found to be P21 and the resolution of the data extended to 2.5 Å (Table 1). The model 119

was rebuilt by hand using Coot (49) and refined using Refmac (50) with the NCS restraints 120

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 7: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

7

option used during all stages of refinement. Four of the molecules (chains A-D) had 121

significantly better electron density than the other two molecules (chains E and F). 122

The crystals were grown in Tacsimate (Hampton Research) which is predominantly 123

malonic acid and there was unambiguous density for malonic acid in the four protomers 124

with good electron density. Despite the poor density for two of the molecules in the 125

asymmetric unit, the model refined to a Rwork/ Rfree of 22.4%/ 25.9% (see Table 1). The 126

Ramachandran plot (from Coot) shows 94.3% of the residues in the most favorable region, 127

4.3% in the allowed region and 1.5% in the outlier region. 128

129

Small Angle X-ray Scattering (SAXS) - AtzF and AtzF467 were dialyzed overnight into 50 mM 130

Tris buffer pH 7.5, 100 mM NaCl. The same buffer was used as the buffer standard during 131

data collection. A dilution series of AtzF (from 6.8 to 0.2 mg/mL) and AtzF467 (from 13.1 to 132

0.4 mg/mL) was prepared, and scattering data were collected for 1 second using a Pilatus 1 133

M photon counting detector (Dektris) with a sample to detector distance of 1.6 m. Ten 134

replicate images were collected for each sample and averaged, with outlier rejection, to 135

control for radiation damage. Data were measured in a Q range from 0.01 to 0.5 Å-1 and at 136

the highest protein concentration the scattering remained above the noise threshold to the 137

edge of the detector. 138

AtzF (75 µL, 6 mg/mL) and AtzF467 (75 µL, 6 mg/mL) were injected on to a size 139

exclusion column (Wyatt Silica Bead column, 300 Å pore size) that had been pre-140

equilibrated with the PO4/NaCl buffer. The column was developed at 0.2 mL/min, and a 141

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 8: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

8

single peak eluted. The SAXS scattering showed no change over the peak (unpublished 142

observations). 143

The radius of gyration (Rg) and total forward scatter (I(0)) for each concentration in 144

the dilution series was calculated using the autorg routine in PRIMUS (51). Fits to atomic 145

models were performed using CRYSOL (52) and dummu atom models were calculated using 146

DAMMIF (53) and superimposed on the high-resolution models using SUPCOMB (54). 147

Molecular weight was calculated from I(0) (55) together with protein concentration 148

measurements and scattering length density and partial specific volume calculated from the 149

protein sequences using the web application MULCh (56). 150

151

Substrate preparation and kinetic assays - The potassium salt of allophanate was prepared 152

by hydrolyzing ethyl allophanate (Sigma Aldrich, MO) with a five molar excess of 1 M 153

potassium hydroxide at 40°C for three hours (42). Potassium allophanate was precipitated 154

overnight in an ice-cold mixture of five volumes of ethanol and one volume of diethyl ether. 155

The precipitate was separated by filtration and subsequently dried and stored in a 156

desiccator. 157

The synthesis of allophanate was confirmed by 13C NMR (Supplementary Fig. 1) using 158

a Varian INOVA-500 NMR Spectrometer. The substrate was dissolved in 1 M KOH prepared 159

in D2O and 2048 scans were performed in total. The purity of the potassium allophanate 160

produced was assessed using a glutamate dehydrogenase (GDH)-coupled assay. Complete 161

AtzF-mediated hydrolysis of allophanate was performed which was then coupled with GDH-162

based ammonia assay in order to determine the purity of allophanate prepared. The GDH-163

dependent decrease in absorbance at 340 nm was measured using a SpectraMax M2 164

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 9: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

9

spectrophotometer (Molecular Devices, California, USA) to follow the consumption of NADH 165

during the reductive amination of α-ketoglutaric acid to form L-glutamate. The hydrolysis of 166

allophanate by AtzF was assumed to have reached completion as a high concentration of 167

AtzF (~ 1 µM) was added and the reaction components of the GDH-linked ammonia assay 168

were added well in excess of allophanate concentration and the decrease in NADH 169

absorbance plateaued before all of the NADH was consumed. Allophanate prepared by the 170

method described above was found to be 86 % pure. 171

To study the pH-dependence of AtzF activity, kinetic assays were performed in the 172

pH range 7-9.5 using allophanate as a substrate and kcat/KM values were plotted against the 173

pH (Supplementary Fig. 3). The kinetic assays were carried out in buffer comprised of 100 174

mM HEPES (hydroxyethyl-piperazineethane-sulfonic acid) and 100 mM CHES (N-Cyclohexyl-175

2-aminoethanesulfonic acid) adjusted to the desired pH. The combined buffer system was 176

used to keep similar conditions throughout the wide pH range used in these assays. Kinetic 177

assays were performed for AtzF, AtzF467 and AtzF H488A using allophanate as the substrate 178

at pH values ranging from 7.0-9.0 at temperatures of 28°C and 4°C, respectively. To account 179

for the temperature-dependent decrease in AtzF activity, assays conducted at 28°C 180

contained 27 nM of enzyme and assays carried out at 4°C contained 53 nM of enzyme. In 181

addition to the enzyme and substrate, each reaction mixture was comprised of 0.5 mM 182

NADH (Sigma Aldrich), 10 mM α-ketoglutaric acid (Sigma Aldrich), 1.4 μM bovine serum 183

albumin (Sigma Aldrich), 12 μM glutamate dehydrogenase (Sigma Aldrich). The Vmax and 184

Michaelis-Menten constant were obtained by non-linear regression analysis and applying 185

robust fit in GraphPad Prism6 software (GraphPad Software, Inc., CA, USA). 186

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 10: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

10

The substrate (allophanate) and product (N-carboxycarbamate) were also analyzed 187

by LC-MS. Enzyme reactions were comprised of 53.5 mM potassium allophanate in 100 mM 188

ammonium acetate buffer pH 9.0. Reactions contained one of the following: no enzyme, 189

0.48 µM AtzF, 0.48 µM AtzF468 or 0.96 µM AtzF468. The reactions were carried out on ice. LC-190

MS analyses were carried out on an Agilent 6550 iFunnel Q-TOF LC/MS system, with 191

samples introduced by direct injection. The mobile phase was comprised of 10mM 192

ammonium acetate pH 9.0 and acetonitrile. Samples were analysed in negative ion mode 193

and m/z values ranging from 50-150 were scanned. 194

195

The effect of pH on enzyme stability was determined by obtaining residual activities 196

after heating AtzF, AtzF467, AtzF H488A in either pH 7 or pH 9 buffer for 5 minutes in an 197

Eppendorf Mastercycler EP (Eppendorf, Hamburg, Germany). The reaction components and 198

method were the same as those used in kinetic assays and 2.2 pmoles of each protein was 199

used in 100 μl of reaction mixture. 200

201

Identification of cyanuric acid mineralization complex - A culture of Pseudomonas sp. strain 202

ADP was grown at 28°C with cyanuric acid as a sole nitrogen source. The composition of 203

media used is as follows: 26.1 mM Na2HPO4.7H2O, 22 mM KH2PO4, 8.5 mM NaCl, 200 μM 204

MgSO4, 2.9 mM sucrose, 3.4 mM trisodium citrate, 4.4 μM CaCl2, 10 mM cyanuric acid, 20 205

ml vitamin stock and 1 ml of 1000 x stock solution of trace elements per liter of culture 206

medium. The 1,000 x trace elements solution comprised of 34.8 μM ZnSO4.7H2O, 15.2 μM 207

MnCl2.4H2O, 485 μM H3BO3, 84 μM CoCl2.2H2O, 5.9 μM CuCl2.2H2O, 12.4 μM 208

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 11: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

11

Na2MoO4.2H2O, and 8.4 μM NiCl2.6H2O and pH of media was adjusted to 7.3 by addition of 209

sodium hydroxide. The composition of the vitamin stock solution was 5 mg of thiamine-HCl, 210

2 mg of biotin, 2 mg of folic acid, 10 mg of nicotinamide, and 10 mg of pyridoxine-HCl per 211

liter of deionised water and 20 ml of vitamin stock solution was added per liter of culture 212

medium. The culture media was made sterile by filtering through 0.2 μm filters (VacuCap™ 213

60, Pall® Life Sciences, NY, USA) before the media was inoculated with a single colony of 214

Pseudomonas sp. strain ADP that had been grown on nutrient agar plates. 215

The culture was grown until it reached an OD600 of 0.6 – 0.7, after which the cells 216

were harvested by centrifugation at 8,000 x g for 15 minutes. Harvested cells were 217

resuspended in 50 mM HEPES, 100 mM NaCl (pH7.5) and lysed using BugBuster® Protein 218

Extraction Reagent (Novagen, Darmstadt, Germany) as per manufacturer instructions. After 219

cell lysis, the soluble fraction was collected by spinning the lysate at 21,000 x g for 15 220

minutes. The cell-free extract was passed through 0.45 μm syringe driven filters (MILLEX® 221

HV, Cork, Ireland) before it was loaded onto a 130 ml size exclusion column (Superdex 200, 222

GE, Uppsala, Sweden). The activities of AtzD, AtzE and AtzF were assessed in each faction 223

using 68 μM of cyanuric acid and biuret and 560 µM allophanate. The rates of background 224

hydrolysis were subtracted in each case. 225

The fraction in which catalytic activities for AtzD, AtzE and AtzF were detected was 226

resolved by SDS-PAGE (4-20% Tris-HEPES-SDS gels, Thermoscientific, Rockford, USA; 227

Supplementary Fig. 2). Certain zones were excised from the gel corresponding to expected 228

locations of AtzD, AtzE and AtzF. In-gel tryptic digest (with the inclusion of ProteaseMAX™ 229

Surfactant (Promega, Madison, USA)) and tandem mass spectral analysis were performed as 230

previously described using an Agilent Chip Cube system coupled to an Agilent XCD ion trap 231

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 12: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

12

mass spectrometer (57). Mass spectra corresponding to common contaminants, such as the 232

added trypsin and keratin, were identified before the remaining mass spectral data were 233

used to search against a database containing all non-redundant UniProtKB/SwissProt 234

protein sequences from Pseudomonas and the Atz sequences using SpectrumMill software 235

(Agilent Rev A.03.03.084 SR4) with its stringent default ‘autovalidation’ settings. 236

237

Results and Discussion 238

Structure of the amidase domain of AtzF – The X-ray crystal structure for AtzF467 is dimeric 239

(Fig. 3A), consistent with observations from previous studies (30). There are three 240

independent dimers found in the asymmetric unit of the crystal: A-B, C-D and E-F. The A-B 241

and C-D dimers are significantly more ordered with average temperature factors (B-factors) 242

of 30 to 33 Å2. The E-F dimer has weaker density and has average B-factors about double 243

that of the other dimers, 65-75 Å2. 244

The overall structure is similar to those of the K. lactis allophanate hydrolase (AHKl; 245

PDB code 4ISS) and the N-terminus of the G. bethesdensis allophanate hydrolase (AHGb; PDB 246

code 4GYS), and to other Ser-cisSer-Lys hydrolases such as the amidase subunit of the 247

Staphylococcus aureus glutamine amidotransferase (PDB code 2DQN; (48)) used for 248

molecular replacement in this study. 2DQN has an rmsd of 1.8 Å with 16 gaps and less than 249

30% sequence identity to AtzF. The dimer interface buries approximately 1150 Å2 of surface 250

per monomer (Fig. 2). The dimer interface is essentially the same as that found in AHKl (PDB 251

code 4ISS) and AHGb, except that the AtzF structure is missing the C-terminal domain, which 252

forms a separate interface between the two monomers in AHKl (Fig. 3A). 253

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 13: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

13

Superposition of the AtzF and AHKl dimers (Fig. 3B and C) gives an rmsd of 1.6 Å for 254

851 residues (out of 895/1226) with 21 gaps (done with the SSM algorithm implemented in 255

Coot). A comparison with AHGb gives an rmsd of 1.3 Å for 857 residues (out of 895/921) 256

with 15 gaps (Fig 3B and C). The two structures are very similar with only a slight shift in 257

helices 2 and 3 at the N-terminus and a C-terminal extension in AHGb of residues 444 to 462 258

that is not seen in the AtzF structure due to the difference in the position of the truncations 259

of each of the proteins. The only other region of significant deviation between the 260

structures is in a mobile loop region between residues 255 and 268 of AtzF (residues 238-261

247 of AHGb). Notably, the structure used as the molecular replacement model, 2DQN, also 262

possessed similar differences to AtzF as AHGb, in addition to an extended loop (residues 322-263

348 in 2DQN and residues 333-340 in AtzF). The active sites of the AtzF, AHGb and AHKl are 264

also essentially identical with the positions of the amino acids essential for catalysis (Ser165, 265

Ser189 and Lys91) and substrate binding (Tyr320 and Arg328) conserved between the three 266

enzymes (Fig. 3D). 267

SAXS data (Fig. 3A-3F) indicate that purified, full-length AtzF is a homotetramer in 268

solution (Fig. 4A), rather than a hexamer as SEC had previously suggested (Shapir et al., 269

2005b, Balotra et al., 2014). (31, 45). This is evident both from the fit of the data to a model 270

of tetrameric AtzF and from the molecular weight of the complex calculated from the SAXS 271

data (Supplementary Table 1). While there is a reasonable agreement overall between the 272

tetrameric model of AtzF and the SAXS data it can be seen that there is some divergence 273

between the two at low angle and around an oscillation in the data at around 0.1 Å-1. This is 274

most likely explained by some conformational flexibility in the C terminal domains of the 275

complex though modeling this structural feature is beyond the scope of this study. The 276

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 14: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

14

tetramer itself is a ‘cylinder’ of sorts that has dimensions of about 180 x 115 x 60 Å. The 277

length of the long axis of the tetramer (180 Å) is longer than might be expected for a 278

tetramer with the molecular weight of AtzF, and this is possibly the cause of this 279

inconsistency between SEC and the more accurate observations from SAXS. 280

Previously, we had shown that removal of the C-terminus of AtzF resulted in a 48 281

kDa protein (AtzF467) which was kinetically indistinguishable from AtzF (Balotra et al., 2014). 282

However, the removal of the C-terminal domain altered the quaternary structure of AtzF, as 283

indicated by the reduction in apparent molecular weight of the enzyme during SEC from 284

~360 kDa to ~110 kDa (dimeric) for the wild-type and truncated proteins, respectively 285

(Balotra et al., 2014). SAXS of the amidase domain of AtzF (AtzF467) confirmed that the 286

truncated variant is indeed a dimer in solution (Fig. 4B). These observations suggest that 287

the C-terminus of AtzF is required for homotetramer formation, but not homodimer 288

formation. In contrast to the data presented by Fan et al. for the AHKl protein (37), which 289

they showed was a monomer in solution when the C-terminal domain was removed (or 290

disrupted by mutations), we find that the truncated AtzF migrates as a dimer in solution, as 291

shown by both SAXS and chromatographic (SEC) data. As the dimeric interfaces are similar 292

between the two proteins, it isn’t clear why the truncated form of AHKl would behave 293

differently to the truncated AtzF. 294

It is interesting that the kinetic parameters for AtzF characterized here are different 295

from those reported by Shapir et al. in 2005 (35); the kcat and KM values for AtzF reported 296

here are ~4-13 s-1 and ~120-300 µM, respectively (Supplementary Table 2), vs. 16 s-1 and 297

1,500 µM in the previous study (35). While it is unclear why there is such a large difference 298

in KM, it is noteworthy that the lowest substrate concentration tested in the report from 299

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 15: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

15

Shapir et al. (34) was 200 µM (i.e. close to the KM determined in the work reported here). It 300

is also of note that the purity of the allophanate used in the work by Shapir was not 301

reported, and so the reported value of 150 µM for the KM reflects the highest possible KM 302

(assuming 100 % substrate purity). 303

304

There is no catalytic advantage conferred by the C-terminus of AtzF in vitro - The N-terminal 305

domain of allophanate hydrolase deaminates allophanate to produce ammonia and N-306

carboxycarbamate (Fig. 1). The role of the smaller C-terminal domain is less clear, however. 307

Fan et al. (37) suggested that the C-terminal domain of AHKl is catalytic, decarboxylating N-308

carboxycarbamate to form carbamate and carbon dioxide. On the basis of in silico substrate 309

docking and mutagenesis studies, it was proposed that a histidine residue in the C-terminal 310

domain (His492; Fig. 2) acts as a catalytic residue. If also true for AtzF, removal of the C-311

terminal domain or substituting the catalytic histidine for alanine would be expected that 312

the rate of ammonia production by AtzF would be reduced and that the intermediate (N-313

carboxycarbamate) would accumulate during the reaction. 314

We were unable to detect of N-carboxycarbamate accumulation in reactions 315

catalyzed by either the full length AtzF or AtzF467 by LC-MS analysis (Supplementry Fig. 2), 316

which may indicate that the C-terminus of AtzF is not catalytic. However, N-317

carboxycarbamate is highly unstable in an aqueous environment, decaying by spontaneous 318

decarboxylation to form carbamate and carbon dioxide. As the stability of carbamates is 319

known to be temperature and pH-dependent (58), the pH of the AtzF catalyzed reactions 320

was raised to 9 and the temperature reduced to 4 °C. Despite the conditions favoring the 321

accumulation of N-carboxycarbamate, we were unable to detect this product in reactions 322

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 16: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

16

catalyzed by either full length AtzF or AtzF467. Although the absence of the accumulation of 323

N-carboxycarbamate in reactions is not direct evidence for a lack of the hypothesized 324

catalytic activity in the C-terminus, it does suggest that such a function would have little 325

physiological relevance. 326

To be sure that our inability to detect N-carboxycarbamate in the reactions was due 327

to the absence of this reaction intermediate, rather than a flaw in our detection method, 328

we also assessed the rate of ammonia production by full length AtzF or AtzF467, and by an 329

AtzF variant in which the ‘catalytic’ histidine residue had been substituted for an alanine 330

residue. The enzymes were tested at a range of pHs and temperatures, to account for the 331

difference in N-carboxycarbamate stability at different temperatures and pHs. 332

For reactions at 28°C, the full-length AtzF was indistinguishable from the two 333

variants at pH values of 8 and below. However, at pH values above 8, the full-length enzyme 334

had kcat values greater than either AtzF467 or AtzF H488A. The difference for AtzF H488A was 335

slight, but the reduction in activity for AtzF467 was considerable (Fig 5A). For reactions at 4°C, 336

the full-length AtzF was indistinguishable from the two variants at pH values of 7.5 and 337

below, however at pH values above 7.5, the full-length enzymes had kcat values greater than 338

AtzF467, albeit the effect on rate was less extensive that at 28°C. At 4°C, there was no 339

difference in kcat value for AtzF and the AtzF H488A variant (Fig 5B). 340

If the differences in apparent kcat values were solely a result of a reduction in 341

catalytic efficiency, it would be expected that the AtzF H488A and AtzF467 variants would be 342

indistinguishable from each other, which they are not. Moreover, the effect on catalytic 343

rate in the variants is reduced at low temperatures where the stabilization of N-344

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 17: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

17

carboxycarbamate should make the effects more pronounced, not less so. Indeed, the kcat 345

values of AtzF H488A were essentially identical to that of the wild-type AtzF protein at 4 °C. 346

As the pH-dependent differences between full-length AtzF and its variants were 347

more pronounced at room temperature than at 4 °C, it is possible that they are a result of 348

differences in thermal stability of the three enzymes. The residual activities of AtzF, AtzF467 349

and AtzF H488A were therefore tested after incubation for 5 minutes at temperatures 350

between 30°C and 70°C and were assayed at pH 7, where no difference had been observed 351

in apparent kcat for the three enzymes, and pH 9, where differences had been observed in 352

apparent kcat for the three enzymes (Fig 6A and 6B). 353

At pH 7, there was no significant difference in the residual activities recovered for 354

AtzF, AtzF467 and AtzF H488A, but at pH 9, the apparent Tm for AtzF467 was reduced from 355

~44°C to ~41°C and although the apparent Tm for AtzF H488A was ~45°C it was lower than 356

AtzF by nearly 1.5 °C. It is therefore plausible that at least a component of the pH-357

dependent difference in apparent kcat values was a result of a pH-dependent reduction in 358

stability for AtzF when the C-terminal was removed or H488 was mutated to an alanine. 359

Regardless of the rate differences observed at higher than physiological pH (i.e. 6.5 - 360

7.5), we were unable to observe a difference in the rate of ammonia release between the 361

wild-type and variant AtzF enzymes tested in vitro at physiological pH conditions. 362

363

AtzF appears to form a large complex with AtzD and AtzE - Allophanate is unstable 364

and has been reported to decompose rapidly to form urea and carbon dioxide under 365

physiological conditions (Fig. 1; (30)). For both the ureolytic pathway and triazine 366

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 18: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

18

degradation pathway, the spontaneous decomposition of allophanate is undesirable, 367

causing an ATP-consuming futile cycle for the ureolytic enzymes and forming a ‘dead-end’ 368

product in triazine hydrolysis that results in the metabolic loss of two of the three nitrogens 369

of the heterocycle. It is likely that the formation of a urea carboxylase/allophanate 370

hydrolase complex in the ureolytic system protects allophanate from decarboxylation by 371

channelling the substrate from one active site to another. It is reasonable to assume that a 372

similar relationship may exist between AtzF and the biuret hydrolase, AtzE, which prevents 373

the production of urea. As AtzF is a relatively inefficient enzyme, with kcat and KM values of 8 374

s-1 and 120 µM, respectively (Supplementary Table 2), it is likely that nitrogen from the 375

degradation of the triazine ring will be lost to the competing abiotic decarboxylation rather 376

than being channeled through allophanate hydrolase in the absence of the formation of a 377

substrate-channeling complex. 378

Size-exclusion chromatography (SEC) was used to fractionate Pseudomonas sp strain 379

ADP cell-free extract obtained from a culture that had been induced for AtzD, AtzE and AtzF 380

production by incubation in a medium containing cyanuric acid as sole carbon source. AtzD, 381

AtzE and AtzF activities were found to co-elute in a fraction with an estimated molecular 382

weight of ~660 kDa. The activities of these three enzymes were not found in other 383

fractions. This observation is consistent with a complex that consists of four AtzD 384

monomers, four AtzF monomers and two AtzE monomers, albeit deriving the stoichiometry 385

of the three enzymes from SEC alone may be misleading. 386

The SEC fraction with AtzD, AtzE and AtzF activities was resolved by SDS-PAGE, and 387

gel sections taken where the three proteins should migrate (based on molecular weight; 388

Supplementary Fig. 3). LC-MS of the tryptic digestion products from these sections showed 389

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 19: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

19

that peptides were present that were consistent with tryptic fragmentation pattern of AtzD, 390

AtzE and AtzF. The sequence coverage for AtzD, AtzE and AtzF was 21% by 9 unique 391

peptides, 25% by 9 unique peptides and 3% by 2 unique peptides, respectively. 392

Unsurprisingly, other proteins identified in the samples were abundant cellular components 393

of Pseudomonas such as translation elongation factors (GI: 169761912, GI: 169761911 and 394

GI: 169760553), and heat shock proteins (GI: 169760669). When considered along with the 395

enzyme activity data from the SEC, these data support the presence of a large protein 396

complex consisting of AtzD, AtzE and AtzF. 397

Production of ammonia was observed upon addition of cyanuric acid to this fraction, 398

with three moles of ammonia produced for every mole of cyanuric acid added, suggesting 399

that the complex mineralized cyanuric acid. Interestingly, the rate of ammonia production 400

was much greater upon addition of cyanuric acid than it was when biuret (the AtzE 401

substrate) was added (Table 2), which may imply that access to the AtzE active site from 402

solvent may be restricted and passage of substrate from the AtzD active site to the AtzE 403

active site may occur via substrate tunneling or channeling. If the AtzD product (1-404

carboxybiuret, Fig. 1) is protected from extended contact with solvent, it is possible that 405

AtzE is a 1-carboxybiuret hydrolase, rather than a biuret hydrolase. Unfortunately, despite 406

considerable effort, we were unable to produce soluble AtzE in sufficient quantity to 407

reconstruct the complex in vitro for further study, and so were unable to probe this 408

hypothesis further. 409

410

Conclusion 411

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 20: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

20

AtzF is a component of a large (~0.7 megaDalton) cyanuric acid-mineralizing 412

complex, comprised of AtzD, AtzE and AtzF. We postulate that this complex protects the 413

unstable intermediate allophanate from spontaneous decarboxylation under physiological 414

conditions by substrate channeling and/or by precluding water from the reactive centers: 415

strategies used in other enzymes that catalyze water sensitive reactions, such as carbamoyl 416

phosphate synthase and the GatCAB glutamine amidotransferase (GatCAB; Weeks et al., 417

2006, Nakamura et al., 2006). Interestingly, ureolytic allophanate hydrolases also form 418

complexes with urea carboxylase. This suggests that allophanate hydrolases have either 419

evolved to form multi-protein complexes more than once, or that they have ‘switched 420

partners’ over the course of their evolution. 421

The biochemical data that we have presented (accumulation of N-carboxycarbamate 422

and rate of ammonia production by AtzF and its variants) fail to support a physiologically 423

relevant catalytic function for the C-terminus of AtzF. It is possible that the proposed N-424

carboxycarbamate decarboxylase activity is relevant under conditions that stabilize N-425

carboxycarbamate, such as a water-limited environment. These conditions could be present 426

in the cyanuric acid-mineralizing complex, albeit further work will be required to elucidate 427

structure/function relationships within the complex. 428

429

Acknowledgements 430

We thank the beamline scientists of the Australian Synchrotron, and the CSIRO Collaborative 431

Crystallisation Centre (www.csiro.au/C3), Melbourne, Australia. We also thank Mr. Chris 432

Blake (Research School of Chemistry, Australian National University) for his help with the 433

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 21: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

21

NMR studies and Drs Carol Hartley and Robyn Russell (CSIRO) for helpful discussions. The 434

authors do not have a conflict of interest to declare. 435

436

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 22: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

22

References 437

1. Tomlin, C. (ed.). 2006. The Pesticide Manual: A World Compendium, Fourteenth ed. BCPC, 438 Alton (UK). 439

2. Thurman, E. M., and M. T. Meyer. 1996. Herbicide metabolites in surface water and 440 groundwater: Introduction and overview, p. 1-15. In M. T. Meyer and E. M. Thurman (ed.), 441 Herbicide Metabolites in Surface Water and Groundwater, vol. 630. Amer Chemical Soc, 442 Washington. 443

3. Gavrilescu, M. 2005. Fate of pesticides in the environment and its bioremediation. Eng. Life 444 Sci. 5:497-526. 445

4. Hayes, T. B., V. Khoury, A. Narayan, M. Nazir, A. Park, T. Brown, L. Adame, E. Chan, D. 446 Buchholz, T. Stueve, and S. Gallipeau. 2010. Atrazine induces complete feminization and 447 chemical castration in male African clawed frogs (Xenopus laevis). Proc. Natl. Acad. Sci. U. S. 448 A. 107:4612-4617. 449

5. Hayes, T., K. Haston, M. Tsui, A. Hoang, C. Haeffele, and A. Vonk. 2003. Atrazine-induced 450 hermaphroditism at 0.1 ppb in American leopard frogs (Rana pipiens): Laboratory and field 451 evidence. Environ. Health Perspect. 111:568-575. 452

6. Hayes, T. B., A. Collins, M. Lee, M. Mendoza, N. Noriega, A. A. Stuart, and A. Vonk. 2002. 453 Hermaphroditic, demasculinized frogs after exposure to the herbicide atrazine at low 454 ecologically relevant doses. Proc. Natl. Acad. Sci. U. S. A. 99:5476-5480. 455

7. Udikovic-Kolic, N., C. Scott, and F. Martin-Laurent. 2013. Evolution of atrazine-degrading 456 capabilities in the environment. Appl. Microbiol. Biotechnol. 96:1175-1189. 457

8. Russell, R. J., C. Scott, C. J. Jackson, R. Pandey, G. Pandey, M. C. Taylor, C. W. Coppin, J. W. 458 Liu, and J. G. Oakeshott. 2011. The evolution of new enzyme function: lessons from 459 xenobiotic metabolizing bacteria versus insecticide-resistant insects. Evol. Appl. 4:225-248. 460

9. Shapir, N., E. F. Mongodin, M. J. Sadowsky, S. C. Daugherty, K. E. Nelson, and L. P. Wackett. 461 2007. Evolution of catabolic pathways: genomic insights into microbial s-triazine 462 metabolism. J. Bacteriol. 189:674-682. 463

10. Wackett, L. P. 2004. Evolution of enzymes for the metabolism of new chemical inputs into 464 the environment. J. Biol. Chem. 279:41259-41262. 465

11. Noor, S., F. Changey, J. G. Oakeshott, C. Scott, and F. Martin-Laurent. 2014. Ongoing 466 functional evolution of the bacterial atrazine chlorohydrolase AtzA. Biodegradation 25:21-467 30. 468

12. Noor, S., M. C. Taylor, R. J. Russell, L. S. Jermiin, C. J. Jackson, J. G. Oakeshott, and C. Scott. 469 2012. Intramolecular epistasis and the evolution of a new enzymatic function. PLoS One 7. 470

13. Raillard, S., A. Krebber, Y. C. Chen, J. E. Ness, E. Bermudez, R. Trinidad, R. Fullem, C. Davis, 471 M. Welch, J. Seffernick, L. P. Wackett, W. P. C. Stemmer, and J. Minshull. 2001. Novel 472 enzyme activities and functional plasticity revealed by recombining highly homologous 473 enzymes. Chem. Biol. 8:891-898. 474

14. Seffernick, J. L., M. L. de Souza, M. J. Sadowsky, and L. P. Wackett. 2001. Melamine 475 deaminase and atrazine chlorohydrolase: 98 percent identical but functionally different. J. 476 Bacteriol. 183:2405-2410. 477

15. Peat, T. S., S. Balotra, M. Wilding, N. G. French, L. J. Briggs, S. Panjikar, N. Cowieson, J. 478 Newman, and C. Scott. 2013. Cyanuric acid hydrolase: evolutionary innovation by structural 479 concatenation. Mol. Microbiol. 88:1149-1163. 480

16. Lima, D., P. Viana, S. Andre, S. Chelinho, C. Costa, R. Ribeiro, J. P. Sousa, A. M. Fialho, and 481 C. A. Viegas. 2009. Evaluating a bioremediation tool for atrazine contaminated soils in open 482 soil microcosms: The effectiveness of bioaugmentation and biostimulation approaches. 483 Chemosphere 74:187-192. 484

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 23: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

23

17. Scott, C., S. E. Lewis, R. Milla, M. C. Taylor, A. J. W. Rodgers, G. Dumsday, J. E. Brodie, J. G. 485 Oakeshott, and R. J. Russell. 2010. A free-enzyme catalyst for the bioremediation of 486 environmental atrazine contamination. J. Environ. Manage. 91:2075-2078. 487

18. Scott, C., G. Pandey, C. J. Hartley, C. J. Jackson, M. J. Cheesman, M. C. Taylor, R. Pandey, J. 488 L. Khurana, M. Teese, C. W. Coppin, K. M. Weir, R. K. Jain, R. Lal, R. J. Russell, and J. G. 489 Oakeshott. 2008. The enzymatic basis for pesticide bioremediation. Indian J. Microbiol. 490 48:65-79. 491

19. Topp, E., F. Martin-Laurent, A. Hartmann, and G. Soulas. 2004. Bioremediation of atrazine-492 contaminated soil, p. 141-154. In J. J. Gan, P. C. Zhu, S. D. Aust, and A. T. Lemley (ed.), 493 Pesticide Decontamination and Detoxification, vol. 863. Amer Chemical Soc, Washington. 494

20. de Souza, M. L., L. P. Wackett, K. L. Boundymills, R. T. Mandelbaum, and M. J. Sadowsky. 495 1995. Cloning, characterization and expression of a gene region from Pseudomonas sp strain 496 ADP involved in the dechlorination of atrazine. Appl. Environ. Microbiol. 61:3373-3378. 497

21. de Souza, M. L., M. J. Sadowsky, and L. P. Wackett. 1996. Atrazine chlorohydrolase from 498 Pseudomonas sp strain ADP: Gene sequence, enzyme purification, and protein 499 characterization. J. Bacteriol. 178:4894-4900. 500

22. Boundy-Mills, K. L., M. L. deSouza, R. T. Mandelbaum, L. P. Wackett, and M. J. Sadowsky. 501 1997. The atzB gene of Pseudomonas sp strain ADP encodes the second enzyme of a novel 502 atrazine degradation pathway. Appl. Environ. Microbiol. 63:916-923. 503

23. Seffernick, J. L., A. Aleem, J. P. Osborne, G. Johnson, M. J. Sadowsky, and L. P. Wackett. 504 2007. Hydroxyatrazine N-ethylaminohydrolase (AtzB): an amidohydrolase superfamily 505 enzyme catalyzing deamination and dechlorination. J. Bacteriol. 189:6989-6997. 506

24. Sadowsky, M. J., Z. K. Tong, M. de Souza, and L. P. Wackett. 1998. AtzC is a new member of 507 the amidohydrolase protein superfamily and is homologous to other atrazine-metabolizing 508 enzymes. J. Bacteriol. 180:152-158. 509

25. Shapir, N., J. P. Osborne, G. Johnson, M. J. Sadowsky, and L. P. Wackett. 2002. Purification, 510 substrate range, and metal center of AtzC: the N-isopropylammelide aminohydrolase 511 involved in bacterial atrazine metabolism. J. Bacteriol. 184:5376-5384. 512

26. Fruchey, I., N. Shapir, M. J. Sadowsky, and L. P. Wackett. 2003. On the origins of cyanuric 513 acid hydrolase: purification, substrates, and prevalence of AtzD from Pseudomonas sp strain 514 ADP. Appl. Environ. Microbiol. 69:3653-3657. 515

27. Seffernick, J. L., J. S. Erickson, S. M. Cameron, S. Cho, A. G. Dodge, J. E. Richman, M. J. 516 Sadowsky, and L. P. Wackett. 2012. Defining sequence space and reaction products within 517 the cyanuric acid hydrolase (AtzD)/barbiturase protein family. J. Bacteriol. 194:4579-4588. 518

28. Cameron, S. M., K. Durchschein, J. E. Richman, M. J. Sadowsky, and L. P. Wackett. 2011. 519 New family of biuret hydrolases involved in s-triazine ring metabolism. ACS Catal. 1:1075-520 1082. 521

29. Shapir, N., G. Cheng, M. J. Sadowsky, and L. P. Wackett. 2006. Purification and 522 characterization of TrzF: biuret hydrolysis by allophanate hydrolase supports growth. Appl. 523 Environ. Microbiol. 72:2491-2495. 524

30. Cheng, G., N. Shapir, M. J. Sadowsky, and L. P. Wackett. 2005. Allophanate hydrolase, not 525 urease, functions in bacterial cyanuric acid metabolism. Appl. Environ. Microbiol. 71:4437-526 4445. 527

31. Shapir, N., M. J. Sadowsky, and L. P. Wackett. 2005. Purification and characterization of 528 allophanate hydrolase (AtzF) from Pseudomonas sp strain ADP. J. Bacteriol. 187:3731-3738. 529

32. Jackson, C. J., C. W. Coppin, P. D. Carr, A. Aleksandrov, M. Wilding, E. Sugrue, J. Ubels, M. 530 Paks, J. Newman, T. S. Peat, R. J. Russell, M. Field, M. Weik, J. G. Oakeshott, and C. Scott. 531 2014. 300-Fold increase in production of the Zn2+-dependent dechlorinase TrzN in soluble 532 form via apoenzyme stabilization. Appl. Environ. Microbiol. 80:4003-4011. 533

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 24: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

24

33. Seffernick, J. L., E. Reynolds, A. A. Fedorov, E. Fedorov, S. C. Almo, M. J. Sadowsky, and L. 534 P. Wackett. 2010. X-ray structure and mutational analysis of the atrazine chlorohydrolase 535 TrzN. J. Biol. Chem. 285:30606-30614. 536

34. Shapir, N., C. Pedersen, O. Gil, L. Strong, J. Seffernick, M. J. Sadowsky, and L. P. Wackett. 537 2006. TrzN from Arthrobacter aurescens TC1 is a zinc amidohydrolase. J. Bacteriol. 188:5859-538 5864. 539

35. Shapir, N., C. Rosendahl, G. Johnson, M. Andreina, M. J. Sadowsky, and L. P. Wackett. 540 2005. Substrate specificity and colorimetric assay for recombinant TrzN derived from 541 Arthrobacter aurescens TC1. Appl. Environ. Microbiol. 71:2214-2220. 542

36. Mulbry, W. W., H. Zhu, S. M. Nour, and E. Topp. 2002. The triazine hydrolase gene trzN 543 from Nocardioides sp strain C190: cloning and construction of gene-specific primers. FEMS 544 Microbiol. Lett. 206:75-79. 545

37. Fan, C., Z. Li, H. Y. Yin, and S. Xiang. 2013. Structure and function of allophanate hydrolase. 546 J. Biol. Chem. 288:21422-21432. 547

38. Lin, Y., and M. St Maurice. 2013. The structure of allophanate hydrolase from Granulibacter 548 bethesdensis provides insights into substrate specificity in the amidase signature family. 549 Biochemistry 52:690-700. 550

39. Whitney, P. A., T. G. Cooper, and Magasani.B. 1973. Induction of urea carboxylase and 551 allophanate hydrolase in Saccharomyces cerevisiae. J. Biol. Chem. 248:6203-6209. 552

40. Whitney, P. A., and T. G. Cooper. 1972. Urea carboxylase and allophanate hydrolase - 2 553 components of a multienzyme complex in Saccharomyces cerevisiae. Biochem. Biophys. Res. 554 Commun. 49:45-51. 555

41. Whitney, P. A., and T. F. Cooper. 1972. Urea carboxylase and allophanate hydrolase - 2 556 components of adenosine triphosphate-urea amidohydrolase in Saccharomyces cerevisiae. J. 557 Biol. Chem. 247:1349-1353. 558

42. Kanamori, T., N. Kanou, S. Kusakabe, H. Atomi, and T. Imanaka. 2005. Allophanate 559 hydrolase of Oleomonas sagaranensis involved in an ATP-dependent degradation pathway 560 specific to urea. FEMS Microbiol. Lett. 245:61-65. 561

43. Shin, S., Y. S. Yun, H. M. Koo, Y. S. Kim, K. Y. Choi, and B. H. Oh. 2003. Characterization of a 562 novel Ser-cisSer-Lys catalytic triad in comparison with the classical Ser-His-Asp triad. J. Biol. 563 Chem. 278:24937-24943. 564

44. Shin, S., T. H. Lee, N. C. Ha, H. M. Koo, S. Y. Kim, H. S. Lee, Y. S. Kim, and B. H. Oh. 2002. 565 Structure of malonamidase E2 reveals a novel Ser-cisSer-Lys catalytic triad in a new serine 566 hydrolase fold that is prevalent in nature. Embo J. 21:2509-2516. 567

45. Balotra, S., J. Newman, N. G. French, L. J. Briggs, T. S. Peat, and C. Scott. 2014. 568 Crystallization and preliminary X-ray diffraction analysis of the amidase domain of 569 allophanate hydrolase from Pseudomonas sp strain ADP. Acta Crystallogr. F-Struct. Biol. 570 Cryst. Commun. 70:310-315. 571

46. Ho, S. N., H. D. Hunt, R. M. Horton, J. K. Pullen, and L. R. Pease. 1989. Site-directed 572 mutagenesis by overlap extension using the polymerase chain-reaction. Gene 77:51-59. 573

47. McCoy, A. J., R. W. Grosse-Kunstleve, P. D. Adams, M. D. Winn, L. C. Storoni, and R. J. 574 Read. 2007. Phaser crystallographic software. J. Appl. Crystallogr. 40:658-674. 575

48. Nakamura, A., M. Yao, S. Chimnaronk, N. Sakai, and I. Tanaka. 2006. Ammonia channel 576 couples glutaminase with transamidase reactions in GatCAB. Science 312:1954-1958. 577

49. Emsley, P., B. Lohkamp, W. G. Scott, and K. Cowtan. 2010. Features and development of 578 Coot. Acta Crystallogr. Sect. D-Biol. Crystallogr. 66:486-501. 579

50. Murshudov, G. N., P. Skubak, A. A. Lebedev, N. S. Pannu, R. A. Steiner, R. A. Nicholls, M. D. 580 Winn, F. Long, and A. A. Vagin. 2011. REFMAC5 for the refinement of macromolecular 581 crystal structures. Acta Crystallogr. Sect. D-Biol. Crystallogr. 67:355-367. 582

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 25: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

25

51. Konarev, P. V., V. V. Volkov, A. V. Sokolova, M. H. J. Koch, and D. I. Svergun. 2003. PRIMUS: 583 a Windows PC-based system for small-angle scattering data analysis. J. Appl. Crystallogr. 584 36:1277-1282. 585

52. Svergun, D., C. Barberato, and M. H. J. Koch. 1995. CRYSOL - A program to evaluate x-ray 586 solution scattering of biological macromolecules from atomic coordinates. J. Appl. 587 Crystallogr. 28:768-773. 588

53. Franke, D., and D. I. Svergun. 2009. DAMMIF, a program for rapid ab-initio shape 589 determination in small-angle scattering. J. Appl. Crystallogr. 42:342-346. 590

54. Kozin, M. B., and D. I. Svergun. 2001. Automated matching of high- and low-resolution 591 structural models. J. Appl. Crystallogr. 34:33-41. 592

55. Mylonas, E., and D. I. Svergun. 2007. Accuracy of molecular mass determination of proteins 593 in solution by small-angle X-ray scattering. J. Appl. Crystallogr. 40:S245-S249. 594

56. Whitten, A. E., S. Z. Cai, and J. Trewhella. 2008. MULCh: modules for the analysis of small-595 angle neutron contrast variation data from biomolecular assemblies. J. Appl. Crystallogr. 596 41:222-226. 597

57. Campbell, P. M., H. E. Trueman, Q. Zhang, K. Kojima, T. Kameda, and T. D. Sutherland. 598 2014. Cross-linking in the silks of bees, ants and hornets. Insect Biochem. Mol. Biol. 48:40-599 50. 600

58. Frahn, J. L., and J. A. Mills. 1964. Paper ionophoresis of amino compounds: formation of 601 carbabmates and related reactions. Aust. J. Chem. 17:256-260. 602

603 604

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 26: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

26

Figure legends. 605

Figure 1. Mineralization of atrazine in Pseudomonas sp. strain ADP. Atrazine (a) is 606

hydrolyzed by AtzA,releasing chloride, the product (b) is further hydrolyzed by AtzB, 607

releasing ethylamine (c) and subsequent hydrolysis of the resulting product (d) releases 608

isopropylamine (e), generating cyanuric acid (f). AtzD hydrolyzes cyanuric acid to produce 609

the unstable 1-carboxybiuret (g), which decomposes to biuret (h). Biuret is deaminated to 610

yield allophanate (i) by AtzE, which can spontaneously decompose to urea (j) or can be 611

deaminated by AtzF to produce the unstable N-carboxycarbamate (k) which spontaneously 612

decomposes first to carbamate (l) then to ammonia. 613

614

Figure 2. Alignment of the allophanate hydrolases from Pseudomonas sp. strain ADP 615

(AtzF), Kluyveromyces lactis (AHKl) and Granulibacter bethesdensis (AHGb). The amidase 616

domain is underlined (single) and the C-terminal domain is underlined (double). The active 617

site residues of the amidase domain are marked with diamonds and the proposed catalytic 618

histidine residue in the C-terminal domain is marked with triangle. The position where AtzF 619

was truncated to produce AtzF468 is marked with an arrow. The level of conservation is 620

also shown; identical (*), highly similar (:), similar (.). 621

622

Figure 3. The AtzF structure and comparison to related structures. A) A cartoon diagram of 623

the dimer of the AtzF structure with the secondary structure colored by magenta for beta-624

sheet, cyan for alpha-helices and orange for loop structures. AtzF has two main domains- 625

the catalytic domain and a second all alpha-helical domain that forms the dimer interface- 626

this has been highlighted by coloring these helices in green. B) The AtzF structure 627

superposed with PDB structures 4ISS and 4GYS. 4ISS is colored in cyan and has an extra 628

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 27: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

27

domain which extends away from the rest of the molecule; AtzF is colored green and 4GYS is 629

colored magenta. The structures superpose well despite limited sequence identity with 630

rmsd values of 1.3 to 1.6 Å. C) Comparison of the AtzF, 4ISS and 4GYS dimers. The figure 631

shows how the dimers are similar and how the 4ISS extra domain helps the dimer formation 632

for this protein. D) The catalytic site of these proteins. In two cases we see substrate 633

mimetics bound into the catalytic site: in the case of AtzF we see clear density for malonate 634

whereas for the 4ISS structure there is tartrate. 635

636

Figure 4. Analysis of the multimerisation of AtzF and Atzf467 by SAXS. A) high resolution 637

model of an AtzF tetramer (green and blue) superimposed on a dummy atom model derived 638

from the SAXS data (grey density). B) high resolution model of dimeric AtzF467 (green and 639

blue) superimposed on a dummy atom model derived from the SAXS data (grey density). C) 640

The fit calculated using CRYSOL between the SAXS data and the AtzF tetramer model shown 641

in panel A. D) The fit calculated using CRYSOL between the measured AtzF467 protein and a 642

model of tetrameric AtzF467 based on the AtzF tetramer model. E) The fit calculated using 643

CRYSOL between dimeric AtzF and the measured AtzF SAXS data. F) Fit between the dimeric 644

AtzF467 model shown in panel B and the measured AtzF467 SAXS data. 645

646

Figure 5. Activity of AtzF (and fragments) at different pHs and temps. Comparison of kcat 647

values at different pH levels obtained at 28°C (A) and 4°C (B) for AtzF, AtzF467and AtzF 648

H488A. All the data points plotted had standard error less than 5.3% and standard errors are 649

presented in the Supplementary Table 2. 650

651

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 28: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

28

Figure 6. Effect of pH on residual activity of AtzF and its variants. Comparison of 652

differences in residual activity of AtzF, AtzF467 and AtzF H488A at pH 7 (A) and pH 9 (B). 653

Enzymes were incubated in pH 7 or pH 9 buffers for 5 minutes at temperatures ranging from 654

30-70°C before testing for residual activity. 655

656 657

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 29: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

29

Table 1. Data collection and refinement statistics 658

PDB = 4CP8

Data collection

Space group P21

Cell dimensions

a, b, c (Å) 82.45, 179.23, 112.61

α, β, γ (°) 90, 106.63, 90

Resolution (Å) 2.50 (2.64 - 2.50)

Rsym or Rmerge 0.150 (0.810)

I / σI 12.6 (2.8)

Completeness (%) 100 (100)

Redundancy 7.7 (7.7)

Refinement

Resolution (Å) 40.0 – 2.50

No. reflections 102,548

Rwork / Rfree (%) 22.4/ 25.9

No. atoms 20,462

Protein 20,132

Ligand/ion 28

Water 312

B-factors

Protein 44.2

Ligand/ion 30.3

Water 24.2

R.m.s. deviations

Bond lengths (Å) 0.010

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 30: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

30

Bond angles (°) 1.367

*Values in parentheses are for highest-resolution shell. 659 660

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 31: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

31

Table 2. Rates of ammonia release by the SEC fraction containing AtzD, AtzE and AtzF 661

activities. Cyanuric acid, biuret and allophanate were added as substrate. The SEC fraction 662

was estimated to contain proteins with a molecular weight of ~660 kDa. N.D.; not done, due 663

to the low rate of substrate hydrolysis. 664

Substrate Cyanuric acid Biuret Allophanate Substrate:ammonia ratio 3.2 N.D. 1.1

Rate of ammonia release (nM/s) 45.4 1.0 180.0

N.D. the stoichiometry of substrate to ammonia was not determined for biuret, because of the low rate of 665

ammonia production. 666

667

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 32: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

32

668

Table 3. 13C NMR shifts of allophanate and possible impurities. 669

Compound Structure 13C Chemical Shift (ppm)

Reported30

13C Chemical Shift (ppm)

observed

Urea

O

NH2 NH2

162.89 163.45

Bicarbonate

O

O- OH

165.23 169.10

Allophanate NH2

O

O-

NHO

159.36158.37

158.87 157.97

670

671 672 on O

ctober 17, 2020 by guesthttp://aem

.asm.org/

Dow

nloaded from

Page 33: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 34: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 35: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 36: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 37: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from

Page 38: on April 2, 2020 by guest · 2014-10-27 · 100 United States) and cloned into the pETCC2 expression vector (15). The resultant plasmid 101 was used to transform electrocompetent

on October 17, 2020 by guest

http://aem.asm

.org/D

ownloaded from