3-(Bromoacetyl)chloramphenicol, an active site directed inhibitor for chloramphenicol acetyltransferase.

Bacterial resistance to the antibiotic chloramphenicol is normally mediated by chloramphenicol acetyltransferase (CAT), which utilizes acetyl coenzyme A as the acyl donor in the inactivation reaction. 3-(Bromoacetyl)chloramphenicol, an analogue of the acetylated product of the forward reaction catalyzed by CAT, was synthesized as a probe for accessible and reactive nucleophilic groups within the active site. Extremely potent covalent inhibition was observed. Affinity labeling was demonstrated by the protection afforded by chloramphenicol at concentrations approaching Km for the substrate. Inactivation was stoichiometric, 1 mol of the inhibitor covalently bound per mole of enzyme monomer, with complete loss of both the acetylation and hydrolytic activities associated with CAT. N3-(Carboxymethyl)histidine was identified as the only alkylated amino acid, implicating the presence of a unique tautomeric form of a reactive imidazole group at the catalytic center. The proteolytic digestion of CAT modified with 3-(bromo[14C]-acetyl)chloramphenicol yielded three labeled peptide fractions separable by reverse-phase high-pressure liquid chromatography. Each peptide fraction was sequenced by fast atom bombardment mass spectrometry; the labeled peptide in each case was found to span the highly conserved region in the primary structure of CAT, which had been tentatively assigned as the active site. The rapid, stoichiometric, and specific alkylation of His-189, taken together with the high degree of conservation of the adjacent amino acid residues, strongly suggests a central role for His-189 in the catalytic mechanism of CAT.

[1]  W. V. Shaw,et al.  Comparison of chloramphenicol acetyltransferase variants in staphylococci. Purification, inhibitor studies and N-terminal sequences. , 1979, The Biochemical journal.

[2]  W. V. Shaw,et al.  Affinity and hydrophobic chromatography of three variants of chloramphenicol acetyltransferases specified by R factors in Escherichia coli , 1976, FEBS letters.

[3]  Y. Nitahara,et al.  Kinetic studies on enzymatic acetylation of chloramphenicol in Streptococcus faecalis , 1979, Antimicrobial Agents and Chemotherapy.

[4]  H. Meloche Bromopyruvate inactivation of 2-keto-3-deoxy-6-phosphogluconic aldolase. I. Kinetic evidence for active site specificity. , 1967, Biochemistry.

[5]  G W Taylor,et al.  FAB-mapping of recombinant-DNA protein products. , 1983, Biochemical and biophysical research communications.

[6]  W. V. Shaw,et al.  Chloramphenicol Acetyltransferases Specified by fi− R Factors , 1973, Antimicrobial Agents and Chemotherapy.

[7]  W V Shaw,et al.  Mechanism of R factor-mediated chloramphenicol resistance , 1968, Journal of bacteriology.

[8]  L. Hersh Kinetic studies of the choline acetyltransferase reaction using isotope exchange at equilibrium. , 1982, Journal of Biological Chemistry.

[9]  W. V. Shaw,et al.  The reactivity of sulfhydryl groups at the active site of an F-factor--specified variant of chloramphenicol acetyltransferase. , 1978, European journal of biochemistry.

[10]  J. Robinson,et al.  Inhibition of pyruvate dehydrogenase multienzyme complex from Escherichia coli with a radiolabeled bifunctional arsenoxide: evidence for an essential histidine residue at the active site of lipoamide dehydrogenase. , 1984, Biochemistry.

[11]  B. S. Hartley,et al.  Primary structure of a chloramphenicol acetyltransferase specified by R plasmids , 1979, Nature.

[12]  R. Kitz,et al.  Esters of methanesulfonic acid as irreversible inhibitors of acetylcholinesterase. , 1962, The Journal of biological chemistry.

[13]  J. Drenth,et al.  Methylation of histidine-48 in pancreatic phospholipase A2. Role of histidine and calcium ion in the catalytic mechanism. , 1980, Biochemistry.

[14]  B. Plapp Application of affinity labeling for studying structure and function of enzymes. , 1982, Methods in enzymology.

[15]  P. Traub,et al.  Physical and functional heterogeneity of ribosomal proteins. , 1966, Journal of molecular biology.

[16]  S. Iida,et al.  The DNA sequence of an IS1‐flanked transposon coding for resistance to chloramphenicol and fusidic acid , 1980, FEBS letters.

[17]  W. V. Shaw,et al.  Use of a synthetic oligonucleotide to identify a chromosomal gene for chloramphenicol acetyltransferase in a plasmid-bearing Flavobacterium. , 1984, Journal of general microbiology.

[18]  K. Wilson,et al.  Comparison of buffers and detection systems for high-pressure liquid chromatography of peptide mixtures. , 1981, The Biochemical journal.

[19]  F. C. Hartman Haloacetol phosphates. Characterization of the active site of rabbit muscle triose phosphate isomerase. , 1971, Biochemistry.

[20]  W. Jencks,et al.  Acetyl-coenzyme A: arylamine N-acetyltransferase. Role of the acetyl-enzyme intermediate and the effects of substituents on the rate. , 1971, The Journal of biological chemistry.

[21]  M. L. Bender,et al.  Methylation of histidine-57 in alpha-chymotrypsin by methyl p-nitrobenzenesulfonate. A new approach to enzyme modification. , 1970, Biochemistry.

[22]  D. Vapnek,et al.  Nucleotide sequence analysis of the chloramphenicol resistance transposon Tn9 , 1979, Nature.

[23]  D. Blow,et al.  Role of a Buried Acid Group in the Mechanism of Action of Chymotrypsin , 1969, Nature.

[24]  W. V. Shaw,et al.  Chloramphenicol Acetyltransferases Determined by R Plasmids from Gram-negative Bacteria , 1978 .

[25]  F. C. Hartman,et al.  Isolation and sequencing of an active-site peptide from Rhodospirillum rubrum ribulosebisphosphate carboxylase/oxygenase after affinity labeling with 2-[(bromoacetyl)amino]pentitol 1,5-bisphosphate. , 1983, Biochemistry.

[26]  W. V. Shaw Chloramphenicol acetyltransferase: enzymology and molecular biology. , 1983, CRC critical reviews in biochemistry.

[27]  W V Shaw,et al.  Analysis of the mechanism of chloramphenicol acetyltransferase by steady-state kinetics. Evidence for a ternary-complex mechanism. , 1984, The Biochemical journal.