Molecular basis of bacterial resistance to organomercurial ... - CiteSeerX

2 downloads 0 Views 1MB Size Report
is quite sluggish in turnover of methyl mercuric chloride at 1 min, n-butylmercuric chloride at 15 min, and phenylmercuric acetate at 15 min'. cis-. Butenylmercuric.
Molecular

basis

of bacterial

organomercurial

and inorganic MARK

CHRISTOPHER

T

AND

L. VERDINE

GREGORY

Department 02115,

WALSH,

D. DISTEFANO,

Chemistry and Molecular

of Biological

Pharmacology,

Bacteria

mediate

organic nontoxic sible for operon, that also

mercuric salts by metabolic conversion to elemental mercury, Hg(0). The genes responmercury resistance are organized in the mer and such operons are often found in plasmids

resistance

to organomercurial

bear drug resistance determinants. three of these mer genes, merR,

and in-

We have merB, and

merA, and have studied their protein products via protein overproduction and purification, and structural and functional characterization. MerR is a metalloregulatory DNA-binding protein that acts as a repressor.of both its own and structural gene transcription in the absence of Hg(II); in addition it acts as a positive effector of structural gene transcription when Hg(II) is present. MerB, organomercury lyase, catalyzes the protonolytic fragmentation of organomercurials to the parent hydrocarbon and Hg(II) by an apparent SE2

mechanism.

MerA,

FAD-containing

enzyme

and

with homology

mercuric

ion

redox-active

reductase,

is an

disulfide-containing

to glutathione

reductase.

It has

evolved the unique catalytic capacity to reduce Hg(II) to Hg(0) and thereby complete the detoxification scheme. WALSH, C. T; DISTEFANO, M. D.; MooRE, M. SHEWCHUK, L. M.; VERDINE, G. L. Molecular basis of bacterial resistance to organomercurial and inorganic mercuric salts. FASEB]. 2: 124-130; 1988.

J.;

Key Words: DNA-binding tox/ication mercuric reductase

protein

.

heavy

metal

de-

mercury resistanceorgano-

mercury lyase

ORGANOMERCURIAL

COMPOUNDS

ARE TOXIC to Organisms

by virtue of their avid ligation to thiol groups in proteins, which, coupled with their lipophiicity, leads to their bioaccumulation in the food chain (1). The predominant natural form is methyl mercury, arising from biomethylation of mercuric ions via methyl B12 in sewage and sediments (2); aryl mercurials are industrial contaminants, e.g., from the paper industry. The most common inorganic mercuric salt is HgS, comprising up to 7% mercury by weight of the ore, red cinnabar. 124

mercuric

MELISSA Harvard

to

J.

Medical

MOORE,

salts

LISA M. SHEWCHUK,

School, Boston,

M#{128}rssachusetts

USA

ABSTRACT

subcloned

resistance

Some 1010 tons of rock (average 80 tg Hg/kg) are weathered globally each year, releasing 800 tons of Hg(II) into the environment; the oceanic reservoir of mercury is estimated at 200 million tons. The dissolved Hg(II) concentration in water typically ranges from 0.03 to 2.0 sg/liter (i.e., up to 10-8 M) and, with this as backdrop, it is not surprising that bacteria have developed resistance mechanisms to detoxify both organomercurial and inorganic mercuric salts. As shown in the biogeochemical cycle for mercury in Fig. 1, bacteria recycle organomercurials back to nontoxic, volatile, elemental Hg(0) by a two-enzyme strategy. The first enzyme, organomercury lyase (merB gene product), catalyzes a protonolytic fragmentation of a carbon-mercury bond to yield, for example, Hg(II) and methane from a methyl mercuric salt. The second enzyme, mercuric ion reductase (merA gene product), takes the Hg(II) ions thus produced and catalyzes a two-electron reduction to yield Hg(0), using NADPH as the electron donor. Both RHgX and HgX2 salts are toxic because of their coordination to protein sulfhydrylgroups, whereas Hg(0) is inert, nontoxic, and lipophilic, and simply volatilizes out of the bacterial cell. Because of the novel chemistry of these two enzymes we have sought to purify and characterize them both structurally and enzymologically to delineate their catalytic mechanisms. The genes that encode these two enzymes and additional mer resistance components are, in general, plasmid encoded, and in some cases are present on transposable elements (e.g., Tn501) within the plasmids. These genes are inducible with low levels of mercury, are widespread in bacterial populations, and often segregate with antibiotic resistance genes as in the Shigella R100 plasmid (3). The most common mer resistance systems, exemplified by Tn501 and R100, are the so-called narrow-spectrum resistance determinants that, lacking the merB gene, confer resistance to Hg(II) salts but not organomercurials. Broad-spectrum resistance to organomercurials in Escherichia coli and Staphylococcus aureas requires a functional merB gene in addition to the other mer genes. The mer genes of both Tn501 and R100 have been sequenced (4-6) as has the merB gene for Staphylococcus (7).These genes are clustered physically and functionally in an operon; a typical organization is shown in Fig. 2. Each operon contains a regulatory gene, merR, and four or five structural genes. merT

0892-6638/88/0002-0124/$01.50. © FASEB

polymerase-binding shown in Fig. 3B. vents genes

Hg#{176},, 2+ Hg i-

_________

reductase

Figure

1. Biogeochemical

CH5 Hg’

organomercury lyase

cycle of mercury.

Adapted

[Sediment

from

with the RNA polymerase-binding loci. Thus, the dual repressor function of MerR in mediating both negative autoregulation and repression of structural gene expression can be rationalized by consideration of the fact that the overlapping promoter regions for both mer mRNA molecules are occupied by MerR (Fig. 3B). The molecular basis of mer gene transcriptional activation by MerR in the presence of coactivator Hg(II) is currently under examination, but preliminary results indicate that MerR binds to the same site regardless of the presence or absence of the coeffector. Thus it appears that MerR does not necessarily achieve repression by physical occlusion of the RNA polymerasebinding site; furthermore, the activation of RNA polymerase binding to the promotor by MerR appears to be mediated through the same operator site as that which is active in repression. The phenomenon of traninteraction

+

CH3Hg

mercuric

transcription and divergent

sites for the mer structural genes, as This protein-DNA interaction preof the downstream mer structural (upstream) merR gene by a direct

I

ref 3.

and merP encode Hg(II) and RHgX transport and periplasmic binding proteins, respectively (8), whereas merA encodes mercuric ion reductase and merB organomercury lyase. merD, on the other hand, has not yet been shown to encode a translated protein (8). Of these, we have focused on the merR, merB, and merA gene products for subcloning, protein overproduction and purification, and structural and mechanistic characterization. Below, we summarize results on each.

scriptional repression at the same operator

MerR,

and activation being modulated site, as observed in the case

of

is apparently

precedented in an analogous case involving regulation of the arabinose operon by the AraC protein (16). Further studies that are aimed at delineating the nature of the active transcription complex in mer structural gene expression are currently in progress.

MerR:

METALLOREGULATORY

DNA-BINDING

PROTEIN

The effector

MerR

protein is both for the mer structural

a positive and a negative genes and is of interest as

a prototypic sion systems. heat shock

system for metal-regulated gene expresOther examples may include eukaryotic (9) and metallothionein transcriptional regulation (10, 11), as well as iron uptake gene regulation in bacteria (12). As a positive effector, both the MerR protein and low levels of Hg(II) (ca. i0 M) are required to turn on mer structural gene transcription (see ref 13 and refs cited therein). In the absence of Hg(II) the MerR protein acts as a repressor to prevent structural gene transcription and also appears to negatively autoregulate its own synthesis (13, 14); the latter property necessarily dictates that the repressor be present in vanishingly small quantities and thus renders its isolation and purification virtually impossible from host cells bearing the wild-type mer operon. By cloning the merR gene, with its 5’ negative autoregulatory region deleted by Bal-31 exonuclease digestion, behind the tac promoter we recently obtained good overproduction of the MerR protein. We have purified it to homogeneity and verified that it is a 144-residue polypeptide (16 kDa), transcribed divergently from the mer structural genes, and is active as a 32-kDa dimer (15). Specific interaction with the mer promoter/operator DNA region was detected by DNase footprinting studies (Fig. 3A), which showed that the MerR dimer protects a palindromic 26-base-pair (bp) DNA fragment that straddles the -10/-35 RNA BACTERIAL

RESISTANCE TO MERCURIALS

Initial analysis of the primary sequence of the 144residue MerR polypeptide suggests two possible domains (15). The NH2-terminal half has two 20-residue stretches with a propensity for formation of a DNA-binding helix-turn-helix motif at residues 9-29 and 55-75. This supersecondary structural motif is the DNA recognition element in a variety of prokaryotic repressor proteins (see ref 17 for a compilation). The second domain of interest is in the COOH-terminal portion where residues 115-126 contain a C X C H X7 C sequence .

.

.

.

.

0

R

transport protein regulatory protein

Under

study

reductase

merR

-*

regulatory

protein

Controlling

mer

DNA activity -‘

enzyme detoxifying

to -+

Hg” by reduction

Hg#{176}

enzyme cleaving organomercurials to RH and Hg

Figure 2. Typical arrangement of genes in a mer operon conferring broad-spectrum resistance to both organomercurial and inorganic mercuric salts.

125

A 1 2 3 4 56 II

II

I

lism, with the carbon-cobalt bonds in methyland 5’-deoxyadenosyl-coenzyme B12 the best studied (18, 19). Organomercurial bonds, unlike many other carbonmetal bonds, are quite stable in water and are resistant to protonolytic decomposition; less than 1% protonolytic fragmentation of CH3HgC1 to OH4 occurs after treatment with concentrated HC1 for 100 mm. Thus the task of this enzyme is not a trivial one. To facilitate investigation we subcloned the merB gene, encoded on plasmid R831 in E. coli J53-1, behind the phage T7 promotor. Using a double-plasmid, T7 RNA polymerase system (20) we overproduced MerB to 3% of the cell protein. This permitted its ready purification to homogeneity in an active form (21), which is a monomer of 22 kDa without detectable metals or cofactors. Broad substrate specificity for primary, secondary, and tertiary alkyl mercuric halides as well as for allyl, vinyl, and aryl mercuric halides was observed. The enzyme is quite sluggish in turnover of methyl mercuric chloride at 1 min, n-butylmercuric chloride at 15 min, and phenylmercuric acetate at 15 min’. cis-

7 GC I

I

I

I

Butenylmercuric chloride is the fastest substrate yet detected at 240 min5 (22).Although these numbers are not impressive on an absolute scale, the methyl case is

B 26 bp Protected Region

I

I 5 3

mwmRNA

#{149} -CCATATCGCTTGACTCCGTACATGAGTACGGAAGTAAGGTTACGCTATCCAATTT-3’

-GGTATAGCGAACTGAGGCATGTACTCATGCCTTCATTCCAATGCGATAGGTTAAA-5’

m,A mANA

Figure

I protection assay for the interaction of MerR of the Tn501 mer operon. Lanes 1-7 correspond to 0, 2, 50, 500, 0.8, 20, and 200 pM MerR protein. B) Partial structure of the operator/promotor region for the Tn501 meT operon. The 26-bp protected region refers to DNA that is protected from DNase attack in the presence of MerA protein, as shown in A. with

3. A) DNase

the regulatory

region

that is a candidate for an Hg(II)-binding site. Both the NH2and 000H-terminal domains of MerR are being probed by engineering of mutant MerR proteins in which putatively critical residues have been specifically replaced. The mutants will be analyzed directly for DNA-binding properties, metal ion responsiveness, and in vivo resistance phenotype effects. The ultimate goal is to understand precisely how a metal-based molecular switch affects specific gene transcription. MerB:

ORGANOMERCURY

LYASE

io-

108-fold and the aryl mercurial 6 x fold over the nonenzymic rates (23). To probe how the enzyme might cleave carbonmercury bonds, we analyzed the stereochemical outcomes with cis-butenylmercuric chloride and endonorbornyl mercuric bromide (2). Both substrates were processed with retention of configuration in D2O, which argues against radical intermediates. A discrete carbonium ion seems unlikely because that should add water to produce an alcohol product, not the observed alkane. A third possibility, a discrete carbanion, seems too high in energy (consider CH3HgC1 proceeding via a naked CH3 species) to be rapidly produced, suggesting that a concerted SE2 mechanism of carbon-mercury cleavage and carbon-hydrogen formation, as proposed for nonenzymic protonolysis of vinyl and alkyl mercurials (24-26), is occurring in the enzyme’s active site. This would be the first case of an enzymic SE2 process. A mechanistic scheme is proposed in Fig. 4 where coordination by two enzyme cysteine SH groups is hypotheaccelerated

-

sRHgSR

CII

H

RH)

..ii

4 -

Organomercury

that it utilizes Carbon-metal 126

lyase,

MerB,

an organometallic bonds are rare

is an unusual enzyme in species as a substrate. participants in metabo-

A

RSH

>

S-H

Excess

Figure 4. A mechanistic proposal cury lyase by an SE2 mechanism.

+

RSHQSR

A-H for catalytic

action

of organomer-

WALSH

ET AL.

sized to activate the carbon-mercury bond for protonolysis (23) because such nucleophiic acceleration of electrophilic reactions at carbon-mercury bonds (e.g., with 12) are known (27). Cleavage via the four-center transition state shown would yield the product hydrocarbon, R-H, and enzyme bis-coordinated Hg(II). Only with exogenous thiols present in the buffer will Hg(II) exchange occur, leaving the enzyme ready for another catalytic cycle (22).

MerA:

MERCURIC

ION

the following

Men

product,

exhibits

Hg(II)

to Hg(0)

+

NADP

+

NADPH

+

H

-‘

Hg(0)

+

reductases

N jj

Tn21

p1258

:

sequenced

to

date,

whereas

AS

two

C

II

I

76

-

MerA

8

119

leu gly

-

27

-

82

lys pro gly glu val

-

met thr

ala ala

pro

-

37

-

met thr

asp ser

ala

-

14

134

-

thr

val asn val gly

555

-

gIn leu ser Icysicysi

ala gly

-

-

140 561

_IJ_l_

pairs

in the

proteins

of the

Tn501 mer operon.

pairs

of mercuric

reductase,

we have

initiated

a muta-

genesis/protein engineering approach (31, 32, 34). We have generated the single ala and ser and double ala mutants listed in Table 1, and have isolated the wildtype and each mutant enzyme in 100-mg quantities via a tac promoter-based overproduction scheme coupled with Orange A matrix affinity chromatography. The mutants were assessed for both in vivo Hg(II) resistance metric,

ties

phenotype and in vitro and catalytic properties.

physical,

include

spectrophoto-

The catalytic properNADPH oxidation,

Hg(II)-dependent oxidase activit)c NADPH-thioNADP transhydrogenation, and 5,5 ‘-dithiobis(2-nitrobenzoate) reduction as well as anaerobic 203Hg(II) volatilization [via reduction to Hg(0)]. Anaerobic reoxidation of reduced enzyme-bound fiavin by Hg(II), which probes a single NADPH-dependent

catalytic

event,

was

also

examined.

Among the active site mutants, ser135 cys140 and ala135 cys140 retain ca. 1-2% levels of the wild-type Hg(II) reductase activity whereas the cys135 ser140 and cys135 ala140 mutants are 5-10 times lower still. Thus both cys135 and cys14o are clearly of catalytic consequence in Hg(II) reduction with the ala135 ala140 double mutant having only 1/2300 the rate of Enz-FADH2 reoxidation by Hg(II) of that calculated for the wild type. On the other hand, all active site mutants do show high

ss aa

levels

471 aa

Figure 5. Primary sequence positioning of cysteines (small black bars) in various mercuric reductases and human glutathione reductase. Alignment of the active site cysteine pairs (AS) shows that Tn501, Tn21, and p1258 mercuric reductases have two additional conserved pairs of cysteine residues (N and C) not found in human glutathione reductase (see refs 4, 5, 32a).

BACTERIAL RESISTANCE TO MERCURIALS

Cysteine

561

547 aa

liii

6.

cysteine pair of MerA, to its active site for reduction. A variant of this “bucket brigade” scheme is displayed in Fig. 7. To assess the respective roles of the three cysteine

of transhydrogenase activity, some up to sixfold wild type, which suggests that at least the NADPH-FAD half-reaction is unimpaired. The ala135 cysj40 and ser135 cys140 enzymes colorfully signal their mutant constitution for they are green and red, respectively, rather than the yellow of wild-type enzyme. This optical perturbation is regiospecific because cys135 R140 enzyme species are consistently yellow. The interpretation, in agreement with the glutathione reductase X-ray structure (35), is of a cys14o thiolateFAD charge transfer complex as shown in Fig. 8. The enzyme color of the green and red mutant enzymes returns to that of unperturbed FAD yellow as the cys14o

that

ii

Human Glutathione Reductase

ala sen ala Icysicysi

-

2RSH

unpaired cysteines are not (Fig. 5). Collected in Fig. 6 are all the cysteine pairs found in the Tn501 MerR, P, T, and A gene products. Given the avidity of Hg(II) for bis-thiol ligation, Brown (33) has proposed a scheme of sequential “handing over” of Hg(II) from one thiol pair to the next, starting with MerP, to MerT, to the 10,13

Tn501

-

ala

stoichiometry:

Most enzymological studies have been conducted with the Tn501-encoded enzyme (28, 29) for which the DNA sequence is available (5). Initial studies revealed that this FAD-containing enzyme possesses a redox-active site disulfide spanning residues 135-140. In this region, mercuric reductase displays remarkable active site homology with the fiavin-containing disulfide oxidoreductases, glutathione reductase and lipoamide dehydrogenase (30). However, only mercuric reductase can generate Hg(0) catalytically with a turnover number of ca. 400 min. In recent work we have begun to investigate what structural features contribute to this unique activity (31, 32). In addition to the cysteine pair at positions 135 and 140, Tn501 mercuric reductase has two more pairs of cysteines at residues 10,13 and 558,559 (there being a total of 561 residues). These pairs are conserved in all mercuric

21

31

Figure

Hg(SR)i

Ieu val

MerP

ala

Ehis

-

113

REDUCTASE

Mercuric ion reductase, the merA gene the unique catalytic ability to reduce

with

MerR

of the

127

merR

merT

merP

merA

merD

wwwwwwwwwwwwwww’

I

I

I

I

I

functionl

I

unknown

r.gulator protein (XerR)

Hg#{176}

Cytoplasm

transport protein O(.rT)

Periplasm

0 scavenger protein (NerP)

Figure reductase

7. Proposed (MerA)

scheme

for sequestration

by a bucket

brigade

mechanism

of Hg(II) ions in the periplasm involving the paired cysteine

thiolate anion 5-6, reflecting

is titrated. This thiol exhibits a pKa of the 2-3 pKa unit stabilization of the thiolate anion in the charge transfer complex. Analysis of the NH2- and COOH-terminal cysteine pairs by protein engineering reveals quite different outcomes from mutagenesis of these residues. The cysteine at 10 and/or 13 can be changed to alanine without de-

and their sequential passage to the active site of mercuric residues of MerP, MerT, and MerA. Adapted from ref 33.

tectable effect on the in vivo mercury resistance phenotype. Thus this pair is dispensable, and indeed an 86-501 proteolytic fragment (29) of wild-type MerA is fully active in vitro. The 000H-terminal cysteines at 558,559, however, are required because the ala558 ala559 mutant is inactive for Hg(II) reduction (but fully active in transhydrogenation) in vitro and imparts an HgC12

10-

E U

/ C

9.2 8E

C 5,

0

U

#{149} 7.

0

UI.

6-

C 0

C 0

o

5-

U C

I...-

0

0

U

‘C

w

4-

3

C 0 U C

I.1 /

a,

E

.,-

4

5

I

I

6

7

8

9

10

pH 3-

w 2I0I 300

I 400

I 500

Wavelength Figure

8. Dissipation

of the cys140 thiolate

128

of the charge transfer pertubation, in the ala,,5 cys,40 mutant

anion

resulting mercuric

I 600

(nm)

in the change ion

700

of enzyme

color from red to yellow, by reversible

titration

reductase.

WALSH ET AL.

TABLE 1. Current known MerA mutations whose catalytic and structural properties have been examined

4.

P.; A.;

BARRINEAU,

SUMMERS,

mercury

NH,

terminus

Redox-active

COOH

Mutant

Native

Region

disulfide

cys,o cys,,

1) cys10 ala,, 2) ala10 ala,,

cys,,,

3) ala,,5 4) ser,,, 5) cys,,, 6) cys1,, 7) ala13,

cys140

8) ala,,8 ala,,g

cys,,8 cys,59

terminus

cys,40 cys,40 ala,40 ser140 ala140

supersensitive phenotype (31, 32) in vivo to cells carrying this merA gene mutation. Whether one or both of these COOH-terminal cysteine residues functions to shuttle Hg(II) into the active site, and/or provide tn- or tetradentate coordination to Hg(II) along with cys135 and cys140 at the moment of Hg(II) reduction, remains to be determined. CONCLUSIONS Bacteria

display

to many toxic heavy metal ions, including silver, gold, mercury, chromium, arsenic, nickel, and cadmium. Of these the molecular basis of mercury resistance is now the best understood and the toxicity of mercury to higher life forms the most publicized. Dissection of the

often

plasmid

encoded,

bacterial strategy for metabolism of organomercurial and inorganic Hg(II) salts has turned up three remarkable proteins from this corner of bioinorganic toxicol-

ogy. The MerR protein is a metalloregulatory DNAbinding protein that acts as a metal-specific negative and positive switch for transcription of the mer genes. The merB gene product, organomercury lyase, is a novel catalyst working to effect protonolytic fragmentation of a carbon-metal bond by an apparently concerted four-center SE2 transition state. It also shows protonolytic activity for a few organotin compounds (23). Mercuric ion reductase, the merA gene product, is a newly discovered member of the class of redox-active disulfidecontaining flavoproteins uniquely evolved to reduce Hg(II) to Hg(0). We acknowledge financial support from the National Institutes of Health, General Medical Institute. Also, we thank our coworkers Drs. Au, Begley, Fox, O’Halloran, Schultz, and WaIts for their past contributions, acknowledged in specific references, to the mer research in this group, which was conducted in the biology and chemistry departments at MIT

REFERENCES

197: 329-332; 1977. T. Plasmid-determined

resistance

FOSTER,

bial

drugs

ion reductase

poson the

operon

and

toxic

Rev. 47: 361-409;

metal

ions

in bacteria.

1983.

BACTERIAL RESISTANCE TO MERCURIALS

to antimicroMicrobiol.

of

the

Inc

Fil

plasmid

operons

Tn501:

of the plasmid

the beginning

R100

and

of the operon

trans-

including

regulatory

region and the first two structural Acad. Sci. USA 81: 5975-5979; 1984. 7. GRIFFIN, H.; FOSTER, T.; SILVER, S.; Misit&, T. Cloning and DNA sequence of the mercuricand organomercurial-resistance determinants of plasmid pDU 1358. Proc. Nati. Acad. Sci. USA 84: 3112-3116; 1987. genes. Proc. Nati.

8.

A. Organization, expression, and evolution of genes for mercury resistance. Annu. Rev. Microbiol. 40: 607-634; 1986. 9. Wu, B.; KINGSTON, R.; M0RIM0T0, R. Human HSP7O SUMMERS,

promoter contains at least two distinct regulatory domains. Proc. Nail. Acad. Sci. USA 83: 629-633; 1986. 10. YAGLE, M.; PALMITER, R. Coordinate regulation of mouse metallothionein I and II genes by heavy metals

11.

D.

HAMER,

Mol. Cell. Biol. 5: 291-294; 1986. Annu. Rev. Biochem. 55:

Metallothionein.

913-951;

1986.

BAGG, A.;

NEILANDS,

J. Mapping of a mutation affecting regulation of iron uptake systems in Escherichia coli K-12.J. Bacleriol. 161: 450-453; 1985. 13. FOSTER, T.; GINNITY, F. Some mercurial resistance

12.

plasmids

from

different

merR regulatory the mer operon 14.

incompatibility

groups

specify

functions

that both repress and induce of plasmid R100. J. Bacteriol. 162:

773-776; 1985. P.; FORD, S.;

LUND,

BROWN,

N. Transcriptional

regula-

tion

of the mercury-resistance genes of transposon Tn501. J. Gen. Microbiol. 132: 465-480; 1986. 15. O’HALLORAN, T.; WALSH, C. Metalloregulatory DNAbinding protein encoded by the merR gene: isolation and

16.

characterization.

R. The

SCHLEIF,

Science 235:

211-214;

L-arabinose operon. coli and Salmonella

1987.

Neidhardt,

F.,

ed. Escherichia typhimurium: cellular and molecular biology, vol. 2. Washington, DC: Am. Soc. Microbiol.; 1987: 1473-1481. 17. PABO, C.; SAUER, R. Protein-DNA recognition. Annu. Rev. Bioc/zem. 53: 293-321; 1984. 18. HALPERN, Mechanisms of coenzyme B 12-dependent rearrangements. Science 227: 869-875; 1985. 19. BABIOR, B., ED. Cobalamins. New York: Wiley; 1975. 20. TABOR, S.; RICHARDSON, C. A bacteriophage T7 RNA polymerase/promoter system for controlled exclusive expression of specific genes. Proc. Natl. Acad. Sci. USA

J.

21.

82: 1075-1078; 1985. T.; WALTS, A.; WALSH,

BEGLEY,

mercurial terization. 22.

1. MCAULIFFE, C. Chemistry of mercury. London: Macmillan; 1977: 261-279. 2. RIDLEY, W. P.; DIzIKEs, L.J.; Woon,J. M. Biomethylation of toxic elements in the environment. Science

3.

resistance

W.; JONES, C.; sequence of the

NR1.J. Mol. Appi. Genet. 6: 601-619; 1985. 5. Btowr’i, N.; FORD, S.; PRIDMORE, R.; FRITZINGER, D. Nucleotide sequence of a gene from the Pseudomonas transposon Tn501 encoding mercuric reductase. Biochemistry 22: 4089-4095; 1983. 6. MIs, T.; BROWN, N.; FRITZINGER, D.; PRIDMORE, R.; BARNES, W.; HABERSTROH, L.; SILVER, S. Mercuric

and glucocorticoids. resistances,

P.; JACKSON, S. The DNA

GILBERT, WILSON,

23.

BEGLEY,

C. Bacterial

organo-

lyase: overproduction, isolation and characBiochemistry 25: 7186-7192; 1986.

T.;

WALTS,

A.;

WALSH,

C. Mechanistic

studies

of a protonolytic organomercurial cleaving enzyme: bacterial organomercurial lyase. Biochemistry 25: 7192-7200; 1986. WALSH, C.; BEGLEY, T.; WALTS, A. Bacterial organomercurial lyase: mechanistic studies on a protonolytic organomercurial cleaving enzyme in mercurial detoxification. Bartmann, W.; Sharpless, K. B., eds. Stereochemistry of organic and bioorganic transformations. Wein-

heim,

FRG:

VCH

Publishers;

1986: 73-83.

129

24.

M. Acid cleavage of methylmercunic iodide. Am. Chem. Soc. 79: 5927-5930; 1957. KREEvOY, M.; HANSON, R. The reaction of alkylmercunic iodides with non-halogen acid.J. Am. Chem. Soc.

reductase.

KREEvOY,

J.

25.

83: 626-630;

26.

27.

1961. G.; SAN FILIPPO, J. The mechanism of reduction of alkylmercuric halides by metal hydrides. J. Am. Chem. Soc. 92: 6611-6624; 1970. SAYRE, L.; JENSEN, F. Mechanism in electrophilic ali-

WHITESIDES,

phatic

substitution:

of bromodemercuration

a kinetic and stereochemical study with bromide ion catalysis. J.

29.

Chem. Soc. 101: 6001-6008; 1979. B.; WALSH, C. Mercuric reductase: purification and characterization of a transposon-encoded flavoprotein containing an oxidation-reduction-active disulfide. J. Biol. Chem. 257: 2498-2503; 1982. Fox, B.; WALSH, C. Mercuric reductase: homology to

30.

glutathione reductase and lipoamide dehydrogenase. lodoacetamide alkylationand sequence of active site peptide. Biochemistry22: 4082-4088; 1985. WILLIAMS, C.; ARScorr, L.; SCHULZ, G. Amino acid

Am.

28. Fox,

sequence homology hydrogenase and

130

between human

pig heart esythrocyte

lipoamide deglutathione

Proc. Natl.

Acad.

Sd.

USA

79: 2199-2201;

1982. 31.

WALSH, C.; DISTEFANO, M.; MooRE, M. Catalytic effects of mutagenesis of the paired cysteine residues in

the bacterial enzyme mercuric ion reductase. Oxender, D., ed. Protein structure and design. New York: Liss. In press. 32. WALSH, C.; MoORE, M.; DISTEFANO, M. Conserved cysteine pairs of mercuric ion reductase: an investigation of function via site-directed mutagenesis. Flavins and flavoproteins. Berlin: de Gruyter. In press. 32a LADDAGA, R. A.; CHU, L.; MI5RA, T. K.; SILVER, S. Nucleotide sequence and expression of the mercurialresistance operon from Staphylococcus aureus plasmid pI2S8. Proc. Nail. Acad. Sci. USA 84: 5106-5110; 1987. 33. BROWN, N. Bacterial resistance to mercury-reductio ad absurdum. Trends Biochem. Sci. 10: 400-403; 1985. 34. SCHULTZ, P.; Au, K.; WALSH, C. Directed mutagenesis of the redox-active disulfide in the flavoenzyme mercuric ion reductase. Biochemistry 24: 6840-6848; 1985. 35. SCHULZ, G.; SCHIRMER, R.; SACHSENHEIMER, W.; Piu, E. The structure of the fiavoenzyme glutathione reductase. Nature (London) 273: 120-124; 1978.

WALSH El AL.