• Aucun résultat trouvé

Role of the Arginine 214 in the substrate specificity of OXA-48

N/A
N/A
Protected

Academic year: 2021

Partager "Role of the Arginine 214 in the substrate specificity of OXA-48"

Copied!
29
0
0

Texte intégral

(1)

HAL Id: hal-02566353

https://hal.archives-ouvertes.fr/hal-02566353

Submitted on 7 May 2020

HAL is a multi-disciplinary open access

archive for the deposit and dissemination of sci-entific research documents, whether they are pub-lished or not. The documents may come from teaching and research institutions in France or abroad, or from public or private research centers.

L’archive ouverte pluridisciplinaire HAL, est destinée au dépôt et à la diffusion de documents scientifiques de niveau recherche, publiés ou non, émanant des établissements d’enseignement et de recherche français ou étrangers, des laboratoires publics ou privés.

Role of the Arginine 214 in the substrate specificity of

OXA-48

Saoussen Oueslati, Pascal Retailleau, Ludovic Marchini, Camille Berthault,

Laurent Dortet, Rémy A Bonnin, Bogdan Iorga, Thierry Naas

To cite this version:

Saoussen Oueslati, Pascal Retailleau, Ludovic Marchini, Camille Berthault, Laurent Dortet, et al.. Role of the Arginine 214 in the substrate specificity of OXA-48. Antimicrobial Agents and Chemother-apy, American Society for Microbiology, 2020, 64 (5), pp.e02329-19. �10.1128/aac.02329-19�. �hal-02566353�

(2)

Revised AAC02329-19 1

Role of the Arginine 214 in the substrate specificity of OXA-48 2

3

Saoussen OUESLATI1, Pascal RETAILLEAU2, Ludovic MARCHINI,2 Camille 4

BERTHAULT1, Laurent DORTET1,3,4, Rémy A. BONNIN1,3,Bogdan I. IORGA2, and 5

Thierry NAAS1,3,4* 6

7

1EA7361 “Structure, dynamic, function and expression of broad spectrum -lactamases”,

8

UMR1184, Faculty of Medicine, LabEx LERMIT, Université Paris-Saclay, Le

Kremlin-9

Bicêtre, France

10

2

Institut de Chimie des Substances Naturelles, CNRS UPR 2301, Labex LERMIT, Université

11

Paris-Saclay, Gif-sur-Yvette, France.

12

3

Associated French National Reference Center for Antibiotic Resistance:

Carbapenemase-13

producing Enterobacteriaceae, Le Kremlin-Bicêtre, France

14

4

Bacteriology-Hygiene unit, Assistance Publique/Hôpitaux de Paris, Bicêtre Hospital, Le

15

Kremlin-Bicêtre, France

16

Running title : Arginine 214 and substrate specificity of OXA-48 17

Keywords: oxacillinase, carbapenemase, OXA-232, antibiotic resistance, beta-lactamase, 18

OXA-48-like 19

Word count: Abstract 239; Text: 3996 20

Figures: 2; Tables: 4 21

22

*Corresponding author: Service de Bactériologie-Hygiène, Hôpital Bicêtre, 78 rue du Général 23

Leclerc, 94270 Le Kremlin-Bicêtre, France. Tel : +33 1 45 21 20 19. Fax : +33 1 45 21 63 40. 24

E-mail: thierry.naas@aphp.fr 25

(3)

Abstract 27

Increasing number of variants of the carbapenem-hydrolyzing class D -lactamase 28

OXA-48 are identified in Enterobacterales worldwide. Among them, OXA-181 and OXA-232 29

are of particular interest, as they differ from each other by a single amino-acid (AA) 30

substitution at position 214 (R in OXA-181, and S in OXA-232), that results in reduced 31

carbapenem-hydrolyzing activity for OXA-232. To investigate the role of the AA214, the X-32

ray structure of OXA-232 was determined and the AA214 of OXA-48 and of OXA-232 was 33

replaced by G, L, D, E, S, R, and K using site-directed mutagenesis. These mutants were 34

phenotypically characterized, and three mutants of OXA-232 were purified to study their 35

steady-state kinetic properties. X-ray structure of OXA-232 along with molecular modelling 36

studies showed that the interaction via a salt bridge between R214 and D159 in OXA-48 is 37

not possible with G214 or S214 mutations. In contrast, with K214 that is also positively 38

charged, the interaction with D159 is maintained. With the E214 mutant an alternative 39

binding conformation of imipenem was evidenced that is not compatible with a nucleophilic 40

attack by S70. Thus, imipenem has very poor apparent affinity for the E214 mutant because 41

of its non-productive binding mode. Similarly, we could explain the lack of temocillin 42

hydrolysis by OXA-232, which is due to the unfavorable interaction between the negatively 43

charged R1 substituent of temocillin with the S214 residue. 44

Overall, we demonstrate that the AA214 in OXA-48-like β-lactamases is critical for 45

the carbapenemase activity. 46

47 48

(4)

Introduction 49

The intensive use of antibiotics to treat infections led to the emergence of multidrug 50

resistant pathogens especially in Gram-negative bacteria (GNB). Among -lactams, 51

carbapenems are considered as last resort antibiotics to treat severe infections caused by GNB 52

(1). The major mechanisms of carbapenem resistance in Enterobacterales are (i) the 53

association of a ß-lactamase that very weakly hydrolyze carbapenems (ESBL or a 54

cephalosporinase) with decreased outer-membrane permeability or (ii) inactivation by specific 55

carbapenem-hydrolyzing -lactamases, carbapenemases (1). Genes encoding for these 56

enzymes are mostly harbored by plasmids explaining their rapid spread. -Lactamases are 57

classified into 4 groups (Amber’s classes A to D) based upon sequence homology (1,2). The 58

clinically-relevant carbapenemases in Enterobacterales belong to class A (KPC-type), B 59

(NDM, VIM, IMP) and D (OXA-48-like). OXA-48, initially identified in Turkey (3), has 60

since rapidly spread in the Mediterranean area, middle East, Europe, India and is now turning 61

into a major global threat (4). OXA-48 hydrolyzes penicillins including temocillin, narrow-62

spectrum cephalosporins, and also carbapenems at low rate, but spares expanded-spectrum 63

cephalosporins (ESC) e.g. ceftazidime and cefepime (5). Along with the spread of OXA-48, 64

several variants have been reported that differ from OXA-48 by amino-acid (AA) 65

substitutions or deletions (http://bldb.eu/BLDB.php?class=D#OXA) mostly located in the 5-66

6 loop (6). OXA-181 that differs from OXA-48 by 4 amino-acid substitutions (T103A,

67

N110D, E169Q and S171A) is the second most prevalent OXA-48 variant. OXA-232 that

68

differs from OXA-181 by an additional substitution R214S in the 5-6 loop, is particularly

69

interesting, as its carbapenem-hydrolyzing activity was significantly impaired, as compared to

70

OXA-181 or OXA-48 (7). These results suggested a pivotal role of R214 in the hydrolysis of

71

carbapenems. To further investigate the role of R214 in the hydrolytic profile of OXA-48-like

(5)

carbapenemases, the X-ray structure of OXA-232, and the steady-state kinetic parameters of

73

in-vitro generated OXA-232 mutants at AA position 214 were determined.

74 75

Results 76

Susceptibility testing. To determine the effect of the substitutions at the AA position

77

214 on the hydrolysis of imipenem and temocillin, MIC values of E. coli expressing OXA-48

78

and OXA-232 and their respective point mutant derivatives were determined.

79

The substitution R214S in OXA-48 led to a phenotype similar to that of OXA-232,

80

e.g. reduced MICs for temocillin and imipenem (Table 1). Conversely, OXA-232-S214R re-81

stored MICs for temocillin and imipenem similarly to OXA-48/181. When AA214 is

82

substituted with an uncharged AA, such as G and L, the MIC values were similar to those of

83

the OXA-232. The most interesting results were obtained with substitutions by negatively

84

charged AA at pH 7.0 such as aspartic acid and glutamic acid. Indeed, MIC values for

85

imipenem and temocillin were remarkably affected, similar to E. coli Top10 control isolate.

86

In a second step, we focused on the analysis of mutants of OXA-232:

OXA-232-87

S214K; OXA-232-S214E and OXA-232-S214G, which are representative for each amino

88

acid group: polar positively charged (K), polar negatively charged (E) and uncharged (G).

89

MIC values for other -lactams were performed and compared to those of 232,

OXA-90

181 (OXA-232-S214R) and OXA-48 (Table 2). Overall, MIC values for benzylpenicillin and

91

cephalothin were not affected and MIC values of the mutants for the other carbapenems

92

(meropenem and ertapenem) vary in the same way as for imipenem.

93 94

Biochemical properties determination. To further characterize the impact of the 95

nature of the residue at position 214 on the hydrolytic profile, steady-state kinetic parameters 96

of the 3 OXA-232 mutants (OXA-232-S214G; OXA-232-S214E and OXA-232-S214K) were 97

(6)

determined and compared to those of OXA-48, OXA-181 and OXA-232 (5,7). Overall, the 98

kinetic studies revealed three patterns (Table 3). OXA-232-S214G exhibited hydrolytic 99

activity towards all tested -lactams similarly to those of OXA-232. The mutant OXA-232-100

S214K possessed a higher catalytic efficiency (kcat / Km= 83 mM-1.s-1) for imipenem of 4-101

fold as compared to OXA-232 (kcat / Km= 23 mM-1.s-1), due to a weak increase of the turnover 102

of the enzyme (kcat). The most interesting result was observed with the OXA-232-S214E 103

mutant. Indeed, the substitution by a negatively charged amino acid led to a drastic increase 104

of the Km (>2000 µM) of at least 200-folds as compared to OXA-232 (Km= 9 µM), thus 105

decreasing the catalytic efficiency. Moreover, this mutant totally lost its hydrolytic properties 106

for temocillin. All these results were concordant with the observed MIC values. Taken 107

together, our results confirm that the hydrolysis of imipenem depends on the nature of the 108

residue at position 214. It appears in light of these results that positively charged amino-acids 109

favor hydrolysis of imipenem. 110

111

Crystal Structure of OXA-232. The crystals of OXA-232 were obtained at basic pH 112

in 30% w/v PEG3000 and in the presence of 0.2 M Li2SO4 (Figure 2 A). Glycerol was added 113

to the mother liquor to cryoprotect the crystals before flash-annealing for data collection. 114

They belonged to hexagonal space group, P 62, with 43.8 % solvent as calculated from the 115

Matthews coefficient of 2.19 Å3 Da-1 (8,5). At that time, the best available OXA-48 model in 116

the PDB was 3HBR (1.9Å) (5) obtained from crystals grown at lower pH (7.5 vs 8.5), in PEG 117

4000 then dipped into cryoprotectant ethyleneglycol, and which belonged to the monoclinic 118

space group P 21 with 48.3% solvent (Matthews coefficient, 2.38 Å3 Da-1). After deletion of 119

22 N-ter amino acids, one half of R214S-3HBR coordinates were used as quasi-homologous 120

model for phasing the hexagonal crystal to 2.2 Å resolution. The refined electron density 121

appeared neat throughout the entire backbone and most of side chains of the biological dimer 122

(7)

per asymmetric unit of the crystal. Alternate positions for Q41 and Q53 were observed as well 123

for D230, whose one position permitted a salt bridge with R107 at the dimer interface. The 124

average Real-space correlation factor calculated by Sfcheck software (10) is 0.932. Compared 125

to 3HBR, up to three additional C-ter amino acids could be displayed in chain B of OXA-232 126

(only one in chain A) but not the 9-His-tag chain, disordered in the surrounding solvent. 127

Additionally, some minor discontinuities observed in some parts of 3HBR main chains could 128

be fixed in 5HFO (N50D/K51D, T99C, S150A h-bonding D148A). The rmsd values with the 129

two 3HBR dimers range between 0.3452 and 0.4197 Å (Superpose, (11)). Obviously, the 130

OXA-232 tertiary structure of each monomer features the usual class D β-lactamase fold, with 131

an α-helical region (α3: 73–82; α4: 110-115; α5: 120–130; α6: 132–142; α9: 185–194) and a 132

mixed α-helix/β-sheet region (β1: 26–28; α1: 31–35; β2: 42–48; β3: 53–56; α2: 59–62; β4: 133

196–199; β5: 204–212; β6: 219–227; β7: 232–240). 98.2% of all residues were inside the 134

favored regions of the Ramachandran plot, and 1.8% in the allowed regions. The quaternary 135

structure of OXA-232 is dimeric as observed for OXA-48 around a non-crystallographic two-136

fold axis, in compliance with our data of size-exclusion chromatography. It was already 137

shown that less monomer surface (200Å2) was buried in that dimer formation compared to 138

OXA-10 and it was built upon an intermolecular β-sheet involving β4 from each subunit 139

linked by reciprocal H-bonds between the A199 NH and the A199 carbonyl O of one chain 140

with the corresponding one related by the ncs two-fold axis. Several other H-bonds and salt 141

bridges also participated in the dimeric interface stabilization, and among these interactions, 142

an anionic binding site lying on the ncs two-fold axis is tweezed by two facing R206 from ncs 143

related β5 sheets. This contrasts with the cation-binding site described at the interface of 144

OXA-10, involving other residues replaced in OXA-48. In the original 3HBR structure, a 145

water molecule was placed but on the faith of spherical electron density shape, low B-factor 146

and recurrent presence of halogens in other class D beta-lactamase structures (but not upon 147

(8)

anomalous signal), it was substituted by a residual chloride coming from previous purification 148

steps. The same misinterpretation was also done in in the structure 4S2K (12), where chlorine 149

was also reported in the crystallizing medium composition. If sulfate ions have also been seen 150

making that salt bridge in class D structures grown in the presence of large amount of sulfate 151

salts (e.g. structures 6NLW and 5FDH), here six such anions have been spotted essentially at 152

the molecule surface as well as five opportune glycerol molecules. Carbamylated lysines 153

(KCX 73) can be described in a similar environment in both structures. Regarding the 154

mutation of R214S (the R214 residue was shown in water-mediated interaction with the 155

avibactam drug in the active site in the structure 4S2K (12), this has not structurally disturbed 156

the β5-β6 hairpin and its direct environment, just adding few solvent molecules in the created 157

void (2-3 waters, and one glycerol molecule in chain A) and modifying the L158B and I215B 158

side chain conformers. Deeper into both active sites, respective carbamylated lysine (KCX 159

73) can be described in a similar environment in both structures OXA-232 and OXA-48. We 160

note however the presence of a sulfate ion in interaction with the same residues that recognize 161

the sulfate moiety of the avibactam (Figure 2B). More recently the structure of another 162

variant, OXA-181; which differs from OXA-232 by five substitutions T103A, N110D,

163

E169Q, S171A and R214S, has been released (PDB code 5OE0 (13)) at a similar resolution

164

(2.05 Å). The rmsd with OXA-232 is even lower than the former overlay (0.224 and 0.252Å)

165

in relation to the fact that structure crystallized in the same unit cell dimensions and space

166

group P62 than OXA-232. However, five residue mutations are on the protein surface that

167

could potentially modulate the contacts leading to the crystal growth and crystallization

168

conditions were slightly different with a neutral pH and PEG-mme 5000 instead of PEG 3000

169

or PEG 4000 for 3HBR. Like OXA-232, sulfate has been used in low quantity in the

170

crystallization drop and one anion was reported in the final 5HFO model bound to the amide

171

N of R186A as for OXA-232. Variability in β-lactamase crystal forms remains tenuous to

(9)

explain or even predict, since crystals of OXA-48 in complex with inhibitors grew in the

173

same unit cell dimensions (the choice of P32 space group instead of P62 could be interpreted

174

by the author’s intention to refine two more monomers and active sites independently) in

175

different crystallizing solution conditions (MPD was the major component) and in absence of

176

sulfate.

177 178

Molecular modelling. An in-silico study was performed to identify the structural

179

determinants that could explain the experimentally determined differences between the

180

hydrolytic profiles of OXA-232 and its variants in comparison with OXA-48. The OXA-232

181

structure was used as a starting point and the mutations S214G, S214K and S214E were

182

modeled based on predicted low energy conformations (14). The resulting models showed

183

that lateral chains of serine and lysine mutants are positioned in the same axis and direction as

184

the lateral chain of arginine, without any clashes. The glutamate mutant would be positioned

185

differently, oriented towards the active site, establishing hydrogen bonds with the backbone

186

nitrogen and with the sidechain hydroxyl of T213. This is the only conformation that avoids

187

clashes of this glutamate mutant with the neighboring residues I215 and D159, the latter being

188

involved in OXA-48 in a salt bridge with R214 that maintains a closed conformation of the

189

active site, which would stabilize the substrates (including carbapenems) in a conformation

190

compatible with an efficient hydrolysis, thus influencing the carbapenem turnover (5). For

191

OXA-232 and its variants the salt bridge R214-D159 cannot be formed, and this may explain

192

the decrease of kcat values compared to OXA-48.

193

Molecular docking calculations of imipenem and temocillin on the R214E mutant of

194

OXA-48 provided explanations for the different hydrolytic parameters that were observed

195

experimentally. Imipenem showed two alternative binding modes, one that was known, with a

196

single ionic interaction between the imipenem carboxylate and the sidechain of R250,

(10)

compatible with a nucleophilic attack by S70 (Figure 2C), and a second one, presumably

198

more stable, with two ionic interactions, one between the imipenem carboxylate and the

199

sidechain of R250 as in the previous binding mode, and a second between the positively

200

charged R2 substituent of imipenem and the mutated residue E214 (Figure 2C). This latter

201

binding mode is not productive, as it is not compatible with a nucleophilic attack by S70, thus

202

increasing the apparent Km value. To confirm this hypothesis, we have determined the IC50 for

203

imipenem with OXA-232 and with the mutant OXA-232 S214E, the only variant with an

204

increased Km value. Interestingly, the values obtained (2 µM and 0.059 µM, respectively)

205

showed that the mutant OXA-232 S214E has an IC50 ~34-fold lower as compared to that of

206

OXA-232. On the other hand, temocillin is the only β-lactam tested possessing a negative

207

charge on the R1 substituent. Our docking calculations showed that this negative charge

208

would establish strong unfavorable ionic interactions with the mutated residue E214, thus

209

precluding the binding in a conformation compatible with a nucleophilic attack by S70

210 (Figure 2D). 211 212 Discussion 213

OXA-48-producing Enterobacterales are now endemic in many countries, such as 214

Turkey, Middle East, North Africa, India and have widely spread across Europe. Since the 215

first description of OXA-48, several variants have been described (1,6). These variants can be 216

classified into 3 groups according to their hydrolysis profile. Most of them, including OXA-217

181 or OXA-162 (7), have an enzymatic activity similar to OXA-48 (5). The second group, 218

represented by OXA-163 (7), OXA-247 (15) and OXA-405 (16), have no carbapenemase 219

activity, but instead a marked hydrolytic activity against ESC, similarly to OXA-ESBLs (17). 220

Finally, the third group represented by OXA-244 (18) and OXA-232 (7) exhibit an overall 221

reduced activity towards all ß-lactams including carbapenems as compared to OXA-48. The 222

(11)

amino acid sequence comparison of OXA-48-variants suggested a link between the primary 223

structure and the function of these enzymes. Indeed, all 48-variants with an OXA-224

ESBLs phenotype (loss of carbapenem hydrolysis and the gain of activity towards ESC) have 225

amino-acid deletions in the 5-6 loop (Figure 1). This observation suggests that this loop 226

plays a role in the substrate specificity. The phenotypic study of OXA-244 (which differs 227

from OXA-48 by only one substitution R214G) (18) and the enzymatic study of OXA-232 228

(which differs from OXA-181 by a single substitution R214S) underline that the residue 214 229

is crucial for the carbapenem and temocillin hydrolysis. To confirm this hypothesis, we 230

generated mutants of OXA-48 and OXA-232 at the position 214, and analyzed their 231

hydrolytic profiles including OXA-232-S214R and OXA-48-R214S. Thus, the substitution of 232

S214R for OXA-232 re-stored hydrolytic properties toward temocillin and imipenem. 233

Similarly, the substitution of R214S for OXA-48, led to a drastic decrease of imipenem and 234

temocillin hydrolysis. In order to better understand this phenomenon, 5 mutants at position 235

214 of OXA-48 and OXA-232 were generated. These substitutions corresponded to amino 236

acid that were: polar positively charged (214K), polar negatively charged (214E and 214D), 237

non-polar (214L) and glycine (214G). Overall, the substitutions with D, E, L and G led to a 238

decreased MICs of imipenem and temocillin. The substitution S214K in OXA-232 led to 239

increased MIC values, but still lower than those of 48. The biochemical study of OXA-240

232-S214G and OXA-232-S214K showed that these 2 substitutions did not affect the 241

apparent affinity toward -lactams (Km values were similar to those of 232 and OXA-242

48) but have an effect on the acylation or deacylation steps of the substrate catalysis (kcat 243

values were similar to those of OXA-232 but smaller than those of OXA-48). The catalytic 244

efficiencies (kcat/Km) of these two mutants were in agreement with the MIC values. The most 245

significant differences were observed with OXA-232-S214E, with a drastic decrease in the 246

apparent affinity for imipenem (Km was at least 200 fold higher as compared to OXA-232 247

(12)

and OXA-48) and a loss of temocillin hydrolysis. The glutamate substitution seems to have a 248

direct effect on the apparent affinity of imipenem. Analysis of the 3D structure of OXA-48 249

showed that R214 interacts with D159 via a salt bridge (5), which maintains the shape and the 250

network of water molecules within the binding site. Our molecular modeling study revealed 251

that with the G214 or S214 mutations, this interaction with D159 is lacking, which 252

presumably increases the flexibility of this part of the binding site. In contrast, K is a 253

positively charged polar amino-acid, which appears to maintain the interaction with D159, 254

even with a lateral chain shorter than R. In the mutant E214, we evidenced an alternative 255

binding conformation of imipenem that is not compatible with a nucleophilic attack by S70. 256

In this case, imipenem acts as an inhibitor with an IC50 value of 0.059µM, which explains the 257

significantly higher Km values determined experimentally for this substrate. We also showed 258

that, in the same E214 mutant, the unfavorable interaction between the negatively charged R1 259

substituent of temocillin with the E214 residue precludes the binding of this antibiotic, thus 260

explaining the structural basis responsible for the lack of temocillin hydrolysis observed 261

experimentally. 262

Overall, we demonstrated that the AA position 214 in OXA-48-like β-lactamases is 263

critical for the carbapenemase activity but also for the temocillin hydrolysis. This point is of 264

outmost clinical importance and explains why detection of OXA-244 or OXA-232-producers 265

remains a challenge for clinical microbiology laboratories (18-20). Indeed, these isolates do 266

not grow on ChromID Carba Smart (bioMérieux, Marcy L’Etoile, France), one of the most 267

used type of medium for the screening of CPEs (19). The ChromID Carba Smart is a biplate 268

containing on one side a carbapenem and on the other temocillin, two substrates that are only 269

weakly hydrolyzed by OXA-244 and OXA-232 (19). 270

271

Materials and methods 272

(13)

Bacterial strains. The clinical strain Escherichia coli LIEU (20) expressing the OXA-232

-273

lactamase was used to clone the blaOXA-232 gene. E. coli TOP10 (Invitrogen, Saint-Aubin,

274

France) was used for cloning and mutagenesis experiments and E. coli BL21 DE3 was used

275

for overexpression experiments (Novagen, Fontenay-sous-Bois, France).

276 277

Antimicrobial agents, susceptibility testing and microbiological techniques. 278

Antimicrobial susceptibilities were determined by the disc diffusion technique on Mueller–

279

Hinton agar (Bio-Rad, Marnes-La-Coquette, France) and interpreted according to the

280

EUCAST breakpoints, updated in May 2018 (http://www.eucast.org). MICs were determined

281

using the Etest technique (bioMérieux). Antibiotics were purchased from Sigma (Saint-282

Quentin-Fallavier, France) except temocillin (Eumedica, Brussels, Belgium) 283

284

PCR, cloning, site-directed mutagenesis and DNA sequencing. The recombinant plasmids 285

pTOPO-blaOXA-232 and pTOPO-blaOXA-48, obtained from a previous study (7), were used as a

286

template for the site-directed mutagenesis assays and specific primers were designed for the

287

different mutations, using the program QuickChange Primer Design (Agilent Technologies).

288

QuikChange II Site-Directed Mutagenesis Kit (Agilent Technologies) was used, following the

289

manufacturer’s recommendations, in order to substitute the AA214 into a Glycine (G), a

290

Lysine (K), a Leucine (L), an Aspartic acid (D), a Glutamic acid (E), a Serine (S) and an

291

Arginine (R). Mutagenesis reaction products were transformed in E. coli TOP10 (Invitrogen,

292

Saint-Aubin, France) and selection was performed on TSA plate containing Kanamycin

293

(50g/ml). For the production of OXA-232, genes were amplified by PCR using the forward

294

primer OXA23-256NdeI (5’-aaaaaCATATGaaggaatggcaagaaaacaaa-3’) and the reverse primer

295

OXAXhoI-stop (5’-aaaaaCTCGAGggggaataatttttcctgtttgag-’3) to increase the purification yield.

296

For the mutants OXA-232 S214E, OXA-232 S214G and OXA-232 S214K the forward

(14)

primers OXANdeI (5′-aaaaaCATATGTTGGTGGCATCGATTATCGG-3′) and reverse primer

298

OXAXhoI-stop (5-aaaaaCTCGAGGAGCACTTCTTTTGTGATGGC-3’) were used.

299

Recombinant plasmids were cloned into pET41b (for OXA-232 wt) allowing the expression

300

of the enzyme with an His-tag and pET9a (for OXA-232 mutants) vector (Invitrogen, Life

301

Technologies, Cergy-Pontoise, France), then transformed into E. coli BL21 DE3 (Novagen).

302

All the recombinant plasmids were extracted using the Qiagen miniprep Kit and sequenced

303

using a T7 promoter and M13 reverse primers or T7 terminator (depending on the plasmid)

304

with an automated sequencer (ABI Prism3100; Applied Biosystems). The nucleotide

305

sequences were analyzed using software available at the National Center for Biotechnology

306

Information website (http://www.ncbi.nlm.nih.gov).

307 308

Protein purification. Overnight culture of E. coli BL21 DE3 harboring recombinant 309

plasmids pET41b-OXA-232 or pET9a-OXA-232-mut were used to inoculate 2 L of LB

310

medium broth containing 50 mg/L kanamycin. Bacteria were cultured at 37°C until an OD of

311

0.6 at 600 nm was reached. The -lactamase was induced overnight with 0.2 mM IPTG as

312

inducer and cultures were centrifuged at 6000 g for 15min. The pellets were resuspended with

313

the binding buffer 25mM phosphate sodium pH 7.4, 300 mM K2SO4, 10 mM imidazole for

314

OXA-232 wt and with the buffer 20 mM Bis-Tris H2SO4 (pH 7.2) for mutants of OXA-232.

315

Bacterial cells were disrupted by sonication and the bacterial pellet was removed by two

316

consecutive centrifugation steps at 10 000 g for 1h at 4°C; the supernatant was then

317

centrifuged at 96 000 g for 1h at 4°C. OXA-232 wt was purified with a NTA-Nickel column

318

(GE Healthcare, Freiburg, Germany) by using the elution buffer 25 mM phosphate sodium pH

319

7.4, 300 mM K2SO4, 500 mM imidazole. Mutants of OXA-232 were purified by using 2

320

anion-exchange chromatography HiTrapTM QHP GE Healthcare (20 mM Bis-Tris H2SO4 pH

321

7.2, then 20 mM piperazine H2SO4 pH 9.5) (7). Finally, a gel filtration step was performed for

(15)

the purifications of all -lactamases with 100mM sodium phosphate buffer pH 7 and 150 mM

323

NaCl with a Superdex 75 column (GE Healthcare, Freiburg, Germany). The protein purity

324

was estimated by SDS–PAGE and the pooled fractions were dialyzed against 10 mM

Tris-325

HCl pH 7 for the mutants and against 0.1 M Hepes (pH 7.5) for OXA-232 wt and

326

concentrated using vivaspin 20 (10 000 MWCOPES Sartorius) columns. Protein

327

concentration was determined according to the Bradford method using the Bio-Rad protein

328

assay standard II kit (Bio-Rad, Marnes-la-Coquette, France) with bovine serum albumin

329

(BSA) as a standard.

330 331

Kinetics assays. Steady-state kinetic parameters were determined using a spectrophotometer 332

ULTROSPEC 2000 (Amersham Pharmacia Biotech) and performed at 30°C in 100 mM

333

phosphate buffer (pH 7) for 10 minutes (7). The disappearance of substrate (-lactam

334

antibiotics) was monitored at the specific wavelengths and converted to initial velocities using

335

the specific extinction coefficients. The kcat and Km values were determined by analyzing

-336

lactam hydrolysis under initial-rate conditions using the Eadie-Hoffstee linearization of the

337

Michaelis-Menten equation with SFWIFT II software (7). Fifty percent inhibitory 338

concentrations (IC50) for imipenem were determined in 100 mM sodium phosphate buffer (pH 339

7) and 100 µM of benzylpenicillin as a reporter substrate. 340

341

Crystallization and data collection. Crystal conditions were screened with 10 mg/mL of 342

concentrated protein using a range of commercially available screens. OXA-232 crystals grew

343

in 0.2 M Lithium sulfate, 0.1M Tris pH 8.5 and 30% w/v PEG 3000. Crystals were dipped

344

briefly in cryoprotectant (20% v/v glycerol in the reservoir solution) and flash-frozen in liquid

345

nitrogen prior to be transferred in a stream of nitrogen at 100K (delivered by a Rigaku

X-346

tream cryosystem). A 2.21-Å resolution data set was collected on a Rigaku

(16)

HF rotating-anode generator with Cu Kα radiation, Varimax HF mirror focusing optics

348

(Rigaku) and MAR345 image-plate detector (MAR Research).

349 350

Structure determination and refinement. X-ray diffraction data sets were processed using 351

XDS (21) and AIMLESS (22); the crystals diffracted to 2.2 Å resolution (Table 4). Initial

352

phases were obtained by molecular replacement with MOLREP (23) using OXA-48 (PDB ID

353

3HBR) as a search probe. Refinement was performed by successive and alternate rounds of 354

refinement with BUSTER (24) and model improvement using Coot (25). The final model was 355

evaluated using MolProbity (26). Data collection and refinement statistics are provided in

356

Table 4.

357 358

Molecular modelling. Molecular modelling was performed to evaluate the effects of the 359

mutations on OXA-232 β-lactamase. Substitution of amino acid was performed by mutating

360

the OXA-232 structure (PDB code 5HFO) in silico using the Dunbrack rotamer library

361

(swapaa command) as a part of the UCSF Chimera software (14, 27). The Dunbrack rotamer

362

library predicts the conformation of the amino acid side-chain based on the global energy

363

minimum of the protein. The identification of interatomic clashes based on VDW radii (28)

364

was performed with UCSF Chimera software (14, 27). Three-dimensional structures of the

β-365

lactam ligands were generated using Corina 3.60 (Molecular Networks GmbH, Erlangen,

366

Germany). Molecular docking calculations were performed using Gold (Cambridge

367

Crystallographic Data Centre, Cambridge, UK) (29) and the GoldScore scoring function. The

368

binding site, defined as a 20 Å radius sphere, was centered on the OG oxygen atom of Ser70.

369

All other parameters had default values. The receptor-ligand complex images were produced

370

using UCSF Chimera (14, 27).

(17)

PDB accession codes. The OXA-232 structure atomic coordinates were deposited in the 372

Protein Data Bank with the accession code 5HFO.

373 374

Acknowledgments 375

All the authors are members of the Laboratory of Excellence in Research on 376

Medication and Innovative Therapeutics (LERMIT). This work was partially funded by the 377

University Paris-Saclay, and by grants from the French National Research Agency (ANR-10-378

LABX-33, and ANR-17-ASTR-0018). 379

380

Transparency declarations 381

We have no competing interests to declare. 382

(18)

REFERENCES 384

1. Nordmann P, Naas T, Poirel P. 2011. Global Spread of Carbapenemase-producing

385

Enterobacteriaceae. Emerg Infect Dis 17: 1791–1798.

386

2. Bush K. 2013. The ABCD’s of β-lactamase nomenclature. J Infect Chemother Off J

387

Jpn Soc Chemother 19:549–559.

388

3. Poirel L, Héritier C, Tolün V, Nordmann P. 2004. Emergence of

oxacillinase-389

mediated resistance to imipenem in Klebsiella pneumoniae. Antimicrob Agents

390

Chemother 48:15–22.

391

4. Poirel L, Potron A, Nordmann P. 2012. OXA-48-like carbapenemases: the phantom

392

menace. J Antimicrob Chemother 67:1597–1606.

393

5. Docquier J-D, Calderone V, De Luca F, Benvenuti M, Giuliani F, Bellucci L, Tafi A,

394

Nordmann P, Botta M, Rossolini GM, Mangani S. 2009. Crystal structure of the

395

OXA-48 beta-lactamase reveals mechanistic diversity among class D carbapenemases.

396

Chem Biol 16:540–547.

397

6. Naas T, Oueslati S, Bonnin RA, Dabos ML, Zavala A, Dortet L, Retailleau P, Iorga

398

BI. 2017. Beta-lactamase database (BLDB) – structure and function. J Enzyme Inhib

399

Med Chem 32:917–919.

400

7. Oueslati S, Nordmann P, Poirel L. 2015. Heterogeneous hydrolytic features for

OXA-401

48-like β-lactamases. J Antimicrob Chemother 70:1059–1063.

402

8. Matthews BW. 1968. Solvent content of protein crystals. J Mol Biol 33:491–497.

403

9. Winn MD, Ballard CC, Cowtan KD, Dodson EJ, Emsley P, Evans PR, Keegan RM,

404

Krissinel EB, Leslie AGW, McCoy A, McNicholas SJ, Murshudov GN, Pannu NS,

405

Potterton EA, Powell HR, Read RJ, Vagin A, Wilson KS. 2011. Overview of the

406

CCP4 suite and current developments. Acta Crystallogr D Biol Crystallogr 67:235–

407

242.

(19)

10. Vaguine AA, Richelle J, Wodak SJ. 1999. SFCHECK: a unified set of procedures for

409

evaluating the quality of macromolecular structure-factor data and their agreement

410

with the atomic model. Acta Crystallogr D Biol Crystallogr 55:191–205.

411

11. Krissinel E, Henrick K. 2004. Secondary-structure matching (SSM), a new tool for

412

fast protein structure alignment in three dimensions. Acta Crystallogr D Biol

413

Crystallogr 60:2256–2268.

414

12. King DT, King AM, Lal SM, Wright GD, Strynadka NCJ. 2015. Molecular

415

Mechanism of Avibactam-Mediated β-Lactamase Inhibition. ACS Infect Dis 1:175–

416

184.

417

13. Lund BA, Thomassen AM, Carlsen TJO, Leiros H-KS. 2017. Structure, activity and

418

thermostability investigations of OXA-163, OXA-181 and OXA-245 using

419

biochemical analysis, crystal structures and differential scanning calorimetry analysis.

420

Acta Crystallogr Sect F Struct Biol Commun 73:579–587.

421

14. Dunbrack RL. 2002. Rotamer libraries in the 21st century. Curr Opin Struct Biol

422

12:431–440.

423

15. Gomez S, Pasteran F, Faccone D, Bettiol M, Veliz O, De Belder D, Rapoport M, Gatti

424

B, Petroni A, Corso A. 2013. Intrapatient emergence of OXA-247: a novel

425

carbapenemase found in a patient previously infected with OXA-163-producing

426

Klebsiella pneumoniae. Clin Microbiol Infect Off Publ Eur Soc Clin Microbiol Infect

427

Dis 19:E233-235.

428

16. Dortet L, Oueslati S, Jeannot K, Tandé D, Naas T, Nordmann P. 2015. Genetic and

429

Biochemical Characterization of OXA-405, an OXA-48-Type Extended-Spectrum

β-430

Lactamase without Significant Carbapenemase Activity. Antimicrob Agents

431

Chemother 59:3823–3828.

(20)

17. Poirel L, Naas T, Nordmann P. 2010. Diversity, epidemiology, and genetics of class D

433

beta-lactamases. Antimicrob Agents Chemother 54:24-38.

434

18. Potron A, Poirel L, Dortet L, Nordmann P. 2016. Characterisation of OXA-244, a

435

chromosomally-encoded OXA-48-like β-lactamase from Escherichia coli. Int J

436

Antimicrob Agents 47:102–103.

437

19. Hoyos-Mallecot Y, Naas T, Bonnin RA, Patino R, Glaser P, Fortineau N, Dortet L.

438

2017. OXA-244-Producing Escherichia coli Isolates, a Challenge for Clinical

439

Microbiology Laboratories. Antimicrob Agents Chemother. 61(9).

440

20. Potron A, Rondinaud E, Poirel L, Belmonte O, Boyer S, Camiade S, Nordmann P.

441

2013. Genetic and biochemical characterisation of OXA-232, a

carbapenem-442

hydrolysing class D β-lactamase from Enterobacteriaceae. Int J Antimicrob Agents

443

41:325–329.

444

21. Kabsch W. 2010. XDS. Acta Crystallogr D Biol Crystallogr 66:125–132.

445

22. Evans PR, Murshudov GN. 2013. How good are my data and what is the resolution?

446

Acta Crystallogr D Biol Crystallogr 69:1204–1214.

447

23. Vagin A, Teplyakov A. 2010. Molecular replacement with MOLREP. Acta

448

Crystallogr D Biol Crystallogr 66:22–25.

449

24. Bricogne, G., Blanc, E., Brandl, M., Flensburg, C., Keller, P., Paciorek, W., Roversi,

450

P., Sharff, A., Smart, O. S., Vonrhein, C. & Womack, T. O. BUSTER. Global Phasing

451

Ltd., Cambridge, England 2017.

452

25. Emsley P, Lohkamp B, Scott WG, Cowtan K. 2010. Features and development of

453

Coot. Acta Crystallogr D Biol Crystallogr 66:486–501.

454

26. Williams CJ, Headd JJ, Moriarty NW, Prisant MG, Videau LL, Deis LN, Verma V,

455

Keedy DA, Hintze BJ, Chen VB, Jain S, Lewis SM, Arendall WB, Snoeyink J, Adams

456

PD, Lovell SC, Richardson JS, Richardson DC. 2018. MolProbity: More and better

(21)

reference data for improved all-atom structure validation. Protein Sci Publ Protein Soc

458

27:293–315.

459

27. Pettersen EF, Goddard TD, Huang CC, Couch GS, Greenblatt DM, Meng EC, Ferrin

460

TE. 2004. UCSF Chimera--a visualization system for exploratory research and

461

analysis. J Comput Chem 25:1605–1612.

462

28. Tsai J, Taylor R, Chothia C, Gerstein M. 1999. The packing density in proteins:

463

standard radii and volumes. J Mol Biol 290:253–266.

464

29. Verdonk ML, Cole JC, Hartshorn MJ, Murray CW, Taylor RD. 2003. Improved

465

protein–ligand docking using GOLD. Proteins Struct Funct Bioinforma 52:609–623.

466 467

(22)

Table 1: Minimum inhibitory concentrations (MICs) of -lactams for E. coli Top 10 (pTOPO-OXA-48 and variants), E. coli Top 10

(pTOPO-468

OXA-232 and variants) and E. coli Top 10.

469 470

MIC (mg/L)

β-lactams E. coli Top 10 pTOPO-OXA-48 E. coli Top 10 pTOPO-OXA-232 E. coli

Top 10 wt R214S R214G R214L R214K R214D R214E wt S214R S214G S214L S214K S214D S214E Temocillin >256 32 32 64 128 8 8 32 256 32 64 128 8 8 8 Imipenem 0.75 0.25 0.25 0.25 0.38 0.25 0.25 0.38 0.75 0.25 0.25 0.5 0.25 0.25 0.25 471 472

(23)

473

Table 2: Minimum inhibitory concentrations (MICs) of -lactams for E. coli Top 10 (pTOPO-OXA-48), E. coli Top 10 (pTOPO-OXA-

474

181), E. coli Top 10 (pTOPO-OXA-232 and its variants), and E. coli Top10.

475

MIC (mg/L)

E. coli Top 10 (pTOPO-) E. coli Top 10

β-lactams OXA-232 S214G OXA-232 S214E OXA-232 S214K

OXA-232 OXA-181 OXA-48

Benzypenicillin >256 >256 >256 >256 >256 >256 64 Temocillin 32 8 128 32 >256 >256 8 Cephalothin 16 16 16 16 16 16 4 Imipenem 0.25 0.25 0.5 0.38 0.75 0.75 0.25 Meropenem 0.047 0.032 0.094 0.047 0.125 0.094 0.016 Ertapenem 0.012 0.012 0.094 0.012 0.19 0.094 0.004 476 477

(24)

478

Table 3: Kinetic parameters of OXA-232 mutants

479 480

Substrate

Km (M)a kcat(s-1)a kcat/Km (mM-1.s-1)a

OXA-232

S214E OXA-232 S214G OXA-232 S214K OXA-232 S214E OXA-232 S214G OXA-232 S214K OXA-232 S214E OXA-232 S214G OXA-232 S214K Benzypenicillin 140 ± 5 95± 6.5 95 ± 5 235 ± 20.6 130 ± 11.6 115 ± 9 1678 ± 91.6 1369 ± 166.5 1189 ± 72.2

Temocillin NH 63 ± 6.7 28 ± 3.6 NH 0.01 ± 0.007 0.03 ± 0.007 NH 0.15 ± 0.07 1 ± 0.29

Cephalothin 100±7 133 ± 12.1 160 ± 13.8 8 ± 0.6 4 ± 0.6 5 ± 0.6 81 ± 10.6 30 ± 3.8 31 ±1.7

Imipenem >2000 5 ± 2,6 6 ± 0.7 ND 0.1 ± 0.07 0.5 ± 0.014 ND 20 ± 6.9 83 ± 8.9

NH, no detectable hydrolysis was observed with 1 μM of purified enzyme and 500 μM of substrate; ND, not determined. a Standard deviation 481

values are indicated after the kinetic parameter values. 482

(25)

Table 4: X-ray crystallography and refinement statistics. 483

Data collection

Diffractometer Rigaku RA mm007HF - mar345

Wavelength (Å) 1.5418

Temperature (K) 100

Crystal-to-detector distance (mm) 200

Rotation range per image (°) 1°

Total rotation range (°) 205

Exposure time per image (min) 5

Space group P 62 Cell dimensions a, b, c (Å) 144.08, 144.08, 53.13 α, β, γ (°) 90.00, 90.00, 120.00 Resolution (Å) 17.2-2.21 Rmerge 0.134 (0.527) I/σ(I) 13.8 (4.1) Completeness (%) 97.6 (90.5) Multiplicity 7.0 (5.7) Refinement Resolution range (Å) 17.2-2.21

No. unique reflections 30389

No. of reflections, working set 30389 (2284)

No. of reflections, test set 1534 (102)

Rwork/Rfree 0.172/0.208 (0.343/0.404)

Overall B factor from Wilson plot Wilson B factor (Å2)

24.3

Cruickshank's DPI for coordinate error (Å) 0.179

No. non-hydrogen atoms 4551

Protein 4014

Water 471

Ligand/Ions 66

Average B, all atoms (Å2) 34.72

Protein 33.63

31.01 (m.c) / 36.11 (s.c)

Water 42.85

Ligand/Ions 46.37

(26)

Bond lengths (Å) 0.01

Bond angles (°) 1.03

Ramachandran plot

Most favoured (%) 98.2

Allowed (%) 1.8

Values in parentheses are for the outer shell (2.21-2.36Å) 484

(27)

Legends to figures 486

Figure 1. Amino acid sequence alignment of OXA-48 variants. Asterisks indicate identical 487

residues in all the three sequences, colon indicate a substitution with another amino acid but 488

with the same proprieties. Amino acid motif that are well conserved among class D 489

lactamases are indicated by grey boxes, and the black-outlined box corresponds to the 5-6 490

loop. Positions D159 and R/S214 are indicated by a thin and large arrow, respectively. 491

CLUSTAL O (1.2.4) multiple sequence alignment. 492

493

Figure 2. A Superposition of crystal structures and of 48 (yellow, PDB: 3HBR) and OXA-494

232 (grey; PDB: 5HFO). The 5-6 loop is delimited by the circle. B Partial active-site close-495

up of sulfate 302A-bound OXA-232. Atoms are colored by type (C, white;N, blue; O, red; S, 496

yellow). Hydrogen bonding and electrostatic interactions are depicted as blue dashes. C 497

Alternative docking conformations of imipenem (colored in magenta (A) and yellow (B), 498

closed form of the β-lactam ring) as non-covalent (Michaelis) complexes with the OXA-48 499

R214E mutant (colored in cyan, generated from structure 6P97). Hydrogen bonds and 500

favorable ionic interactions are shown as blue springs. Protein hydrogens are hidden for 501

clarity. D Docking conformation of temocillin (colored in green, closed form of the β-lactam 502

ring) as non-covalent (Michaelis) complex with OXA-48 (colored in pink, PDB 6P97), 503

superposed with the OXA-48 R214E mutant (colored in purple, generated from structure 504

6P97, only Gln214 residue shown). Hydrogen bonds and favorable ionic interactions are 505

represented as blue springs, and unfavorable ionic interactions are represented as yellow 506

thick lines. Protein hydrogens are hidden for clarity. 507

508 509

(28)
(29)

Figure

Table 2: Minimum inhibitory concentrations (MICs) of -lactams for E. coli Top 10 (pTOPO-OXA-48), E
Table 3: Kinetic parameters of OXA-232 mutants 479  480  Substrate  Km (M) a  k cat  (s -1 ) a  k cat /K m  (mM -1 .s -1 ) a  OXA-232     S214E  OXA-232 S214G  OXA-232 S214K  OXA-232    S214E  OXA-232 S214G  OXA-232 S214K  OXA-232      S214E  OXA-232 S214
Table 4: X-ray crystallography and refinement statistics.

Références

Documents relatifs

Therefore, we report here a catalytic living ROMP with a revers- ible CTA (CTA2 or CTA3) that exploits a degenerative reversible chain-transfer polymerization mechanism requiring

Certes, ce serait prendre le risque d’aller droit dans le mur car elle savait bien qu'il était avec la brunette, mais c'était tout de même tentant d'essayer, de le tester avec

Woodhams R, Matsunaga K, Iwabuchi K, Kan S, Hata H, Kuranami M, Watanabe M, Hayakawa K (2005) Diffusion-weighted imaging of malig- nant breast tumors: the usefulness of

Wohlgemuth ( 1957 ) and Büdel ( 1948 ) mentioned that high humidity favours the growth of microorganisms on the combs. In order to monitor the development of such growth, combs

14—The SiC particle distribution, inside the filter media along the filter side wall, resulting from the high magnetic flux density test (Coil #2 at 0.2 T).. The overall direction

The analysis of the field dependence of the measured μSR relaxation rate σ SC ∝ B 2  allows to estimate the ab- solute value of the in-plane magnetic penetration depth λ ab

Comparisons between whole and milled feather exper- iments showed that the percentage of remaining feather was positively correlated with GF/C fragments, proteins, amino acids,

Prognostic impact of amenorrhoea after adjuvant chemotherapy in premenopausal breast cancer patients with axillary node involvement: results of the International Breast Cancer