The Biological Chemistry of Lead

5
223 Recent biophysical studies on the interactions between lead and recombinant proteins and peptides that naturally bind zinc or calcium have provided unparalleled insights into the biological chemistry and molecular toxicology of lead. These studies lay the foundation for the rational design of improved methods for detecting and treating lead poisoning. Addresses Department of Chemistry, Northwestern University, 2145 Sheridan Road, Evanston, IL 60208-3113, USA; e-mail: [email protected] Current Opinion in Chemical Biology 2001, 5:223–227 1367-5931/01/$ — see front matter © 2001 Elsevier Science Ltd. All rights reserved. Abbreviations ALAD δ-aminolevulinic acid dehydratase BLL blood lead level CP-CCCC consensus zinc-binding domain with Cys 4 zinc-binding site CP-CCHC consensus zinc-binding domain with Cys 2 HisCys zinc- binding site CP-CCHH consensus zinc-binding domain with naturally occurring Cys 2 His 2 zinc-binding site HP2 human protamine 2 PKC protein kinase C Introduction Lead poisoning is the one of the most common pediatric health problems in the United States, affecting approxi- mately 890,000 children nationwide at any given time [1]. The sources of this exposure are primarily leaded paint, which was not banned in the United States until 1978, and contaminated soil. Recent studies suggest that much of the existing soil contamination is probably a result of deposi- tion from exhaust from cars that used leaded gasoline, in addition to leaded paint used on the exterior of buildings [2 ]. Lead from these sources exists as or evolves into a variety of Pb 2+ compounds, which are remarkably persis- tent in the environment. Unfortunately, these sources of exposure are often expensive to remediate, and the politics surrounding this issue are complex, suggesting that the legacy of lead poisoning will continue to plague mankind for many years to come [3,4]. Lead poisoning can afflict both children and adults, but the greatest concern is for children, who experience symptoms at significantly lower blood lead levels (BLLs) than do adults [1]. In addition, children tend to develop permanent developmental and neurological problems when exposed chronically to lead, whereas many of the symptoms experi- enced by adults are reversed when exposure is ceased. Although a broad range of epidemiological studies has been conducted on lead poisoning, its molecular underpinnings have remained relatively obscure. However, recent advances in biophysics and molecular biology have provid- ed the tools necessary to study the biological chemistry of lead. These studies have helped to provide insights into the following questions: 1. Out of all of the molecular targets that have been pro- posed for lead, which ones are physiologically relevant? 2. How does lead binding affect the structure and dynam- ics of target proteins? 3. Is lead binding to proteins under thermodynamic or kinetic control? These studies and questions are the subject of this review, which focuses particularly on work from the past two years. Molecular targets for lead Several classes of molecular targets have been proposed to account for the symptoms associated with lead poisoning. With few exceptions, these targets fall into two primary categories: proteins that naturally bind calcium and pro- teins that naturally bind zinc [5–8,9 •• ]. If these interactions are to be physiologically relevant, lead must bind tightly enough to the proposed target to occupy the site(s) under physiologically relevant conditions. To ascertain whether lead binds ‘tightly enough’, both the affinity of lead for the protein and the concentration of free (or ‘bioavailable’) lead must be known. The affinity of lead for a given pro- tein is related to the disocciation constant for the lead–protein complex, . In the simple case, half of the population of target protein (P) will have lead bound when is equal to the concentration of ‘free’ (or ‘bioavail- able’) lead ([Pb]) in the cell. When lead is competing with another metal ion (M, e.g. calcium or zinc) for a binding site, then the relative affinities and concentrations of the two metals must be considered, where: How much bioavailable lead is present in cells? The amount of bioavailable (or ‘free’) lead has not been determined experimentally for most cell types, because of the lack of sensitive and selective fluorescent probes for lead [10 ,11 ]. A child is considered to have lead poisoning if he or she has a total BLL (as measured by atomic absorp- tion spectroscopy) greater than or equal to 10 µg/dl (0.5 µM or 100 parts per billion) [1]. By contrast, a typical person in the United States today who does not have lead poisoning will have a BLL of ~2 µg/dl (0.1 µM or 20 parts per billion). It has been estimated that <5% of the total lead in blood is bound to plasma proteins [12] and that only 0.01% of lead in plasma (~ picomolar lead) is bioavailable [13]. By contrast, the total concentrations of zinc and calci- um in human plasma are about 17 µM and 10 –6 to 10 –3 M, × = M Pb K K MP PbP Pb d M d Pb d K Pb d K The biological chemistry of lead Hilary Arnold Godwin

Transcript of The Biological Chemistry of Lead

Page 1: The Biological Chemistry of Lead

223

Recent biophysical studies on the interactions between leadand recombinant proteins and peptides that naturally bind zincor calcium have provided unparalleled insights into thebiological chemistry and molecular toxicology of lead. Thesestudies lay the foundation for the rational design of improvedmethods for detecting and treating lead poisoning.

AddressesDepartment of Chemistry, Northwestern University, 2145 Sheridan Road,Evanston, IL 60208-3113, USA; e-mail: [email protected]

Current Opinion in Chemical Biology 2001, 5:223–227

1367-5931/01/$ — see front matter© 2001 Elsevier Science Ltd. All rights reserved.

AbbreviationsALAD δ-aminolevulinic acid dehydratase BLL blood lead levelCP-CCCC consensus zinc-binding domain with Cys4 zinc-binding

siteCP-CCHC consensus zinc-binding domain with Cys2HisCys zinc-

binding siteCP-CCHH consensus zinc-binding domain with naturally occurring

Cys2His2 zinc-binding siteHP2 human protamine 2PKC protein kinase C

IntroductionLead poisoning is the one of the most common pediatrichealth problems in the United States, affecting approxi-mately 890,000 children nationwide at any given time [1].The sources of this exposure are primarily leaded paint,which was not banned in the United States until 1978, andcontaminated soil. Recent studies suggest that much of theexisting soil contamination is probably a result of deposi-tion from exhaust from cars that used leaded gasoline, inaddition to leaded paint used on the exterior of buildings[2•]. Lead from these sources exists as or evolves into avariety of Pb2+ compounds, which are remarkably persis-tent in the environment. Unfortunately, these sources ofexposure are often expensive to remediate, and the politicssurrounding this issue are complex, suggesting that thelegacy of lead poisoning will continue to plague mankindfor many years to come [3,4].

Lead poisoning can afflict both children and adults, but thegreatest concern is for children, who experience symptomsat significantly lower blood lead levels (BLLs) than doadults [1]. In addition, children tend to develop permanentdevelopmental and neurological problems when exposedchronically to lead, whereas many of the symptoms experi-enced by adults are reversed when exposure is ceased.Although a broad range of epidemiological studies has beenconducted on lead poisoning, its molecular underpinningshave remained relatively obscure. However, recentadvances in biophysics and molecular biology have provid-ed the tools necessary to study the biological chemistry of

lead. These studies have helped to provide insights into thefollowing questions:

1. Out of all of the molecular targets that have been pro-posed for lead, which ones are physiologically relevant?

2. How does lead binding affect the structure and dynam-ics of target proteins?

3. Is lead binding to proteins under thermodynamic orkinetic control?

These studies and questions are the subject of this review,which focuses particularly on work from the past two years.

Molecular targets for leadSeveral classes of molecular targets have been proposed toaccount for the symptoms associated with lead poisoning.With few exceptions, these targets fall into two primarycategories: proteins that naturally bind calcium and pro-teins that naturally bind zinc [5–8,9••]. If these interactionsare to be physiologically relevant, lead must bind tightlyenough to the proposed target to occupy the site(s) underphysiologically relevant conditions. To ascertain whetherlead binds ‘tightly enough’, both the affinity of lead for theprotein and the concentration of free (or ‘bioavailable’)lead must be known. The affinity of lead for a given pro-tein is related to the disocciation constant for thelead–protein complex, . In the simple case, half of thepopulation of target protein (P) will have lead bound when

is equal to the concentration of ‘free’ (or ‘bioavail-able’) lead ([Pb]) in the cell. When lead is competing withanother metal ion (M, e.g. calcium or zinc) for a bindingsite, then the relative affinities and concentrations of thetwo metals must be considered, where:

How much bioavailable lead is present in cells?The amount of bioavailable (or ‘free’) lead has not beendetermined experimentally for most cell types, because ofthe lack of sensitive and selective fluorescent probes forlead [10•,11•]. A child is considered to have lead poisoningif he or she has a total BLL (as measured by atomic absorp-tion spectroscopy) greater than or equal to 10 µg/dl(0.5 µM or 100 parts per billion) [1]. By contrast, a typicalperson in the United States today who does not have leadpoisoning will have a BLL of ~2 µg/dl (0.1 µM or 20 partsper billion). It has been estimated that <5% of the totallead in blood is bound to plasma proteins [12] and that only0.01% of lead in plasma (~ picomolar lead) is bioavailable[13]. By contrast, the total concentrations of zinc and calci-um in human plasma are about 17 µM and 10–6 to 10–3 M,

×=

MPb

K

KMPPbP

Pbd

Md

PbdK

PbdK

The biological chemistry of leadHilary Arnold Godwin

Page 2: The Biological Chemistry of Lead

respectively. The concentrations of bioavailable zinc andcalcium have been estimated to be 10–12 to 10–6 M and 10–8

to 10–6 M, respectively [13–16]. These estimates suggestthat, in general, lead must bind to proteins with a Kd of≤10–12 M and that lead must bind more tightly than thenative metal ion by at least one to three orders of magni-tude. Although it has been suggested that the localconcentration of lead may be higher in certain organelles(e.g. the nucleus) [17] or specific cell types (e.g. neurons)[12], these requirements serve as a reasonable minimalguideline until more detailed studies on lead biodistribu-tion become available.

Interactions between lead and zinc proteinsThe target for lead that has been studied most thoroughlyin vitro and in vivo is the zinc enzyme δ-aminolevulinicacid dehydratase (ALAD, also called porphobilinogen syn-thase) [8] (Figure 1). ALAD catalyzes the second reactionin the heme biosynthetic pathway, and inhibition of thisenzyme by lead explains (at least in part) the anemia oftenseen in adults and children with high BLLs (≥ 40 µg/dl)[8]. The log of the activity of ALAD in erythrocytesdecreases linearly with the individual’s BLL [18]; leadinhibits ALAD in vitro with an inhibition constant of0.07 pM (versus ) [19]. Why out of all thezinc enzymes in the body is ALAD the only one known tobe inhibited by lead? A recent crystal structure of yeastALAD offers insights into this ‘mystery’. Whereas mostzinc enzymes contain a zinc-binding site with a mixture ofhistidine, cysteine, and carboxylate residues, the yeast andmammalian forms of ALAD contain a unique catalytic

zinc-binding site with three cysteine residues [20,21•].When ALAD is co-crystallized with lead, lead binds prin-cipally to this three-cysteine site (Figure 1). Although thestructure of the lead protein is essentially identical to thatof the zinc form, incorporation of lead non-competitivelyinhibits substrate binding [20]. The suggestion by Erskineet al. [20] that lead prefers the Cys3 site in ALAD becausethis constitutes a tight binding site for lead is borne out byrecent model compound studies: lead binds more tightlythan zinc (about 500 to 1) to a novel tris-thiol ligand(tris(mercaptoarylimdazolyl)borate; Ar = Ph, Mes) but isconsiderably less Lewis-acidic in this environment than iszinc [22••]. These studies provoke several important ques-tions. Does lead bind tightly to other zinc proteins thatcontain three (or more) cysteines in their active sites?Could lead binding to any of these proteins account forsome of the other, more pressing, symptoms associatedwith lead poisoning?

Although no other catalytic zinc sites have been reportedthat contain three or more cysteine residues, there are sev-eral classes of proteins that contain structural zinc-bindingsites with three (Cys3His) or four (Cys4) cysteine residues.These proteins (e.g. retroviral nucleocapsid protein, withCys3His sites, and steroid receptors, with two tandem Cys4sites) act as transcription factors and regulate many of thedevelopmental processes associated with lead poisoning inchildren. Thus, the ability of lead to alter the activity ofthese proteins could account for the pressing developmen-tal problems associated with lead poisoning.

Until recently, the interactions between lead and tran-scription factors had not been probed directly, presumablybecause Pb2+ (electronic structure = [Xe]4f145d106s2) waswidely assumed to be spectrosopically silent. However,when lead binds to cysteine residues in proteins, itexhibits intense lead–thiolate charge-transfer bands in theultraviolet region of the electromagnetic spectrum[23,24,25••,26•] that can provide detailed and quantitativeinformation about lead–protein interactions [25••](Figure 2). Studies on a series of zinc-finger consensuspeptides with different metal-binding sites (consensuszinc-binding domain with naturally occurring Cys2His2zinc-binding site, CP-CCHH; consensus zinc-bindingdomain with Cys2HisCys zinc-binding site, CP-CCHC;and consensus zinc-binding domain with Cys4 zinc-bind-ing site, CP-CCCC) reveal that lead binds tightly tostructural zinc-binding sites ( ) andthat lead binds approximately two orders of magnitudemore tightly than zinc to the Cys4 site [25••]. Furthermore,competition experiments reveal that lead readily displaceszinc from the Cys4 site, which suggests that metal bindingis under thermodynamic, rather than kinetic, control. 1HNMR studies on a Cys2HisCys domain from HIV-nucleo-capsid protein and circular dichroism studies on both thisdomain and CP-CCCC reveal that even though lead bindstightly to these domains, it does not promote proper fold-ing of the site [25••]. These studies lay the foundation for

M10to10 149 −−=PbdK

pM1.6=ZnmK

224 Bio-inorganic chemistry

Figure 1

Lead targets proteins that naturally bind calcium and zinc. Examples ofproteins that are targeted by lead include synaptotagmin, which actsas a calcium sensor in neurotransmission, and ALAD, the secondenzyme in the heme biosynthetic pathway. Each of these proteins wascrystallized with lead as a heavy atom derivative ([8,19,35,36]; SuttonRB, personal communication). These studies revealed that, despite itssize, lead (1.19 Å, blue sphere and circles) can substitute for calcium(0.99 Å, green spheres) in synaptotagmin ([35,36]; Sutton RB,personal communication) and zinc (0.74 Å, red spheres) in ALAD[8,19]. Protein coordinates were obtained from the Protein Data Bank(http://www.rcsb.org/pdb/).

Pb2+ (1.19 Å)Ca2+ (0.99 Å)

Synaptotagmin

Zn2+ (0.74 Å)

ALADCurrent Opinion in Chemical Biology

Page 3: The Biological Chemistry of Lead

detailed molecular studies on the interactions betweenlead and larger proteins that contain structural zinc-bind-ing sites. Parallel studies have recently been reported forthe interactions of lead with the zinc protein human prota-mine 2 (HP2), which plays an important role inspermatogenesis; interactions between lead and HP2 mayaccount for increased infertility rates in men who are occu-pationally exposed to lead. Lead binds to thiol groups inHP2 and alters the structure of the protein [26•]. Theobservation that lead alters the structure of the zinc pro-teins to which it binds is not surprising given thedifferences in coordination preferences (6–8 for Pb2+ ver-sus 4 for Zn2+ and geometries (irregular for Pb2+ versustetrahedral for Zn2+) in structural zinc-binding sites for thetwo metals [27]. However, these results are extremelyimportant because they suggest that lead binding to struc-tural zinc-binding domains should disrupt theDNA-binding activity of the proteins and transcription fac-tors in which they are found.

The effects of lead on the DNA-binding activity ofCys2His2 zinc-finger proteins such as TFIIIA were recent-ly studied using DNase I protection assays. These studiesreveal that when sufficiently high levels of lead are present(5–20 µM), then TFIIIA does not bind DNA [28•] (see alsoUpdate). In addition, recent studies by Goldstein, Bressler,and co-workers [29•] reveal that lead alters expression ofimmediate early genes in PC12 cells, in a protein kinase C(PKC)-dependent fashion. Whether the alterations in geneexpression observed arise from interaction of Pb2+ with a

transcription factor upstream of PKC or solely by directactivation of PKC (see below) is not known. These studiespoint to the need for not only more detailed biophysicalstudies on a wide range of proposed targets, but also forcomprehensive, systemic studies on the mechanisms bywhich lead alters signal transduction in cells.

Interactions between lead and calcium proteinsIn addition to causing developmental problems, lead poison-ing results in pervasive neurological problems in bothchildren and adults [30]. Lead interferes with the ability ofcalcium to trigger exocytosis of neurotransmitters in neuronalcells [31], suggesting that lead might generally target proteinsinvolved in calcium-mediated signal transduction [6]. Thishypothesis was bolstered by the observation by Markovac andGoldstein [13] that picomolar concentrations of lead can acti-vate calcium-dependent PKC. By contrast, EF-hand calciumproteins, such as calmodulin, can be activated by lead, butonly at micromolar lead concentrations [32]. What is the sig-nificance of lead being able to activate PKC? Members of thePKC family regulate many cellular events, ranging from reg-ulation of cell growth to learning and memory [33].Calcium-dependent isoforms of PKC (α, β, and γ) contain aninteresting calcium-binding domain, termed a C2 domain[34]. The C2 domain of PKC contains a multinuclear calci-um-binding site on an exposed loop and binds phospholipidsin a calcium-dependent fashion. Lead promotes phospholipidbinding at lower concentrations than does calcium, suggestingthat lead binds to the calcium site of the C2 domain and thatit binds more tightly than does calcium [13].

The biological chemistry of lead Godwin 225

Figure 2

Lead binding to structural zinc-bindingdomains can be determined directly andquantitatively by monitoring lead–thiolatecharge-transfer bands in the ultraviolet [25•• ].By conducting competition experiments withzinc, the relative affinities of lead and zincwere determined for a series of consensuszinc finger peptides with different metal-binding residues. These studies revealed thatlead binds more tightly than zinc to Cys4sites and that lead can displace zinc fromthese sites on a physiologically relevanttimescale [25•• ]. Adapted from reference[25•• ] with permission.

S N

NSM

(KbPb)

(KbZn)

Wavelength (nm)

250 300 350 4000

2000

4000

6000

8000

10,000

12,000

14,000

16,000

Mol

ar a

bsor

ptiv

ity (

M–1

cm

–1)

Pb(CP-CCHH)

Pb(CP-CCHC)

Pb(CP-CCCC)

Pb2+ in bis-Tris

0.04 43= 0.1

~ <<S N

S SM

S

S S

SM

Page 4: The Biological Chemistry of Lead

Is the effect of lead on C2-containing proteins a general one?Interestingly, two C2 domains are also found in synaptotag-min (syt), the protein that senses calcium influx at nervetermini and mediates calcium-triggered release of neuro-transmitters. The structure of the first C2 domain from syt(Figure 1) was solved using lead as a heavy-atom derivative,and the lead was found to occupy one of the calcium-bind-ing sites ([35,36]; Sutton RB, personal communication).Recent studies by Bouton et al. [37••] reveal that lead bindstightly to syt and that nanomolar concentrations of lead pro-mote binding of synaptotagmin to phospholipid micelles.However, lead interferes with the ability of synaptotagminto bind to its protein partner syntaxin; this inhibition offersa possible molecular explanation for the ability of lead tointerfere with calcium-triggered neurotransmitter release.

Conclusions and remaining questionsOver the past five years, a quantum leap has been made inour understanding of the molecular mechanism of lead poi-soning. Detailed biophysical studies have revealed thatlead binds tightly to both zinc and calcium sites in proteinsand alters their activity. However, lead binds to the ‘best’(cysteine-rich) zinc sites many orders of magnitude moretightly than to the ‘best’ (C2 domain) calcium sites. Thistempts the chemist to say that effects of lead on zinc pro-teins are ‘more important’ than those of lead on calciumproteins. However, the biology of lead is more complex:the multitude of symptoms associated with lead poisoningsuggest that no single target (or class of targets) will explainall of lead’s effects. Additionally, until more is known aboutthe distribution and speciation of lead within the body, wecannot definitively know which proteins are targeted andwhen. Questions that remain to be addressed include, butare not limited to:

1. How is lead taken up into cells?

2. What is the speciation of lead within different cell types andhow is this speciation affected by normal cellular homeostasis?

3. Does lead target Cys4 zinc sites (e.g. in steroid receptors)in vivo?

4. How can lead bind tightly to the calcium sites of synap-totagmin and yet prevent binding to syntaxin?

5. Are other C2 domain proteins affected by lead?

Interestingly, many of the targets that have been explored todate on a molecular level are involved in signal transductionpathways. The implications of these effects are immense:the alteration of one protein molecule by lead could beamplified throughout the subsequent pathway. New toolssuch as DNA microrrays [38] and proteomics [39] are pro-viding the means to test these molecular hypotheses bystudying the effects of lead (and other toxins) on signalingpathways. In addition, recent developments in our under-standing of the mechanism of metal ion transport

[40–44,45•,46•] lay the foundation for understanding howtoxic metals are taken up and distributed about the body.Finally, new developments in fluorescent sensors for metalions offer the promise that we will be able to correlate thesesystematic changes with changes in lead concentrations andspeciation [10•,11•].

UpdateRecent work has demonstrated not only that Pb2+-TFIIIAdoes not bind to DNA, but also that addition of Pb2+ to the Zn2+-TFIIIA-DNA adduct causes dissociation of the protein from DNA [47].

AcknowledgementsThe original experimental work described herein that was conducted byHAG and co-workers was supported by the Burroughs Wellcome Fund (NewInvestigator Award in Toxicology to HAG) and the National Institutes ofHealth (R01 GM58183). HAG is a recipient of a Camille and Henry DreyfusNew Faculty Award, a National Science Foundation CAREER Award, aCamille Dreyfus Teacher-Scholar Award, and a Sloan Research Fellowship.

References and recommended readingPapers of particular interest, published within the annual period of review,have been highlighted as:

• of special interest••of outstanding interest

1. Centers for Disease Control and Prevention: Screening YoungChildren for Lead Poisoning: Guidance for State and Local PublicHealth Officials. Atlanta: USA Department of Health and HumanServices, Public Health Service; 1997.

2. Mielke HW: Lead in the inner cities. Am Sci 1999, • 87:62-73.This paper reviews sources of lead poisoning in the environment; particularemphasis is placed on the recent observations that lead in soil may be amajor contributor to childhood exposure to lead.

3. Lanphear BP: The paradox of lead poisoning and prevention.Science 1998, 281:1617-1618.

4. Needleman HL: Childhood lead poisoning: the promise andabandonment of primary prevention. Am J Pub Health 1998,88:1871-1877.

5. Sunderman FJ, Barber A: Finger-loops, oncogenes, and metals.Ann Clin Lab Sci 1988, 18:267-288.

6. Goldstein GW: Evidence that lead acts as a calcium substitute insecond messenger metabolism. Neurotoxicology 1993, 14:97-102.

7. Simons TJB: Lead-calcium interactions in cellular lead toxicity.Neurotoxicology 1994, 14:77-86.

8. Warren MJ, Cooper JB, Wood SP, Shoolingin-Jordon PM: Leadpoisoning, haem synthesis and 5-aminolaevulinic aciddehydratase. Trends Biochem Sci 1998, 23:217-221.

9. Bressler J, Kim KA, Chakraborti T, Goldstein G: Molecular•• mechanisms of lead neurotoxicity. Neurochem Res 1999,

24:595-600.This paper reviews effects of lead on protein kinase C.

10. Deo S, Godwin HA: A selective, ratiometric fluorescent sensor for• Pb2+. J Am Chem Soc 2000, 122:174-175.The first report of a fluorescent sensor for lead that is ratiometric and bindslead selectively over other metal ions.

11. Li J, Lu Y: A highly sensitive and selective catalytic DNA biosensor• for lead ions. J Am Chem Soc 2000, 122:10466-10467.Catalytic DNA is used as a biosensor for lead.

12. Christensen JM, Kristiansen J: Lead. In Handbook on Metals inClinical and Analytical Chemistry. Edited by Seller HG, Sigel A,Sigel H. New York: Marcel Dekker Inc.; 1994:425-440.

13. Markovac J, Goldstein GW: Picomolar concentrations of leadstimulate brain protein kinase C. Nature 1988, 334:71-73.

226 Bio-inorganic chemistry

Page 5: The Biological Chemistry of Lead

14. Sensi SL, Canzoniero LM, Yu SP, Ying HS, Koh JY, Kerchner GA,Choi DW: Measurement of intracellular free zinc in living corticalneurons: routes of entry. J Neurosci 1997, 17:9554-9564.

15. Fraústo da Silva JJR, Williams RJP: The Biological Chemistry of theElements: The Inorganic Chemistry of Life. Oxford: Clarendon Press; 1991.

16. Thiel EC, Raymond KN: Transition-metal storage, transport, andbiomineralization. In Bioinorganic Chemistry. Edited by Bertini I,Gray HB, Lippard SJ, Valentine JS. Mill Valley, CA: University ScienceBooks; 1994:1-36.

17. Hitzfield B, Taylor DM: Characteristics of lead adaption in a ratkidney cell line. I. Uptake and subcellular and subnucleardistribution of lead. Mol Toxicol 1989, 2:151-162.

18. Millar JA, Battinitini V, Cumming RLC, Carswell F, Goldberg A: Leadand δδ-aminolevulinic acid dehydratase levels in mentally retardedchildren and in lead-poisoned suckling rats. The Lancet 1970,2:695-698.

19. Simons TJB: The affinity of human erythrocyte porphobilinogensynthase for Zn2+ and Pb2+. Eur J Biochem 1995, 234:178-183.

20. Erskine PT, Senior N, Awan S, Lambert R, Lewis G, Tickle IJ,Sarwar M, Spencer P, Thomas P, Warren MJ et al.: X-ray structure of5-aminolaevulinate dehydratase, a hybrid aldolase. Nat Struct Biol1997, 4:1025-1031.

21. Jaffe EK, Martins J, Li J, Kervinen J, Roland L, Dunbrack J: The• molecular mechanism of lead inhibition of human

porphobilinogen synthase. J Biol Chem 2001, 276:1531-1537.Mutagenesis and kinetic data demonstrate that the Cys3 site in ALAD isrequired for catalytic activity and is the site of lead inhibition. Lead can alsobind to a hybrid site involving both ligands from the Cys3 (‘ZnB’) site and lig-ands from a second (‘ZnA’) site.

22. Bridgewater BM, Parkin G: Lead poisoning and the inactivation of•• 5-aminolevulinate dehydratase as modeled by the tris(2-

mercapto-1-phenylimidazolyl)hydroborato lead complex,{[TmPh]Pb}[ClO4]. J Am Chem Soc 2000, 122:7140-7141.

A synthetic analog of ALAD is used to study the replacement of zinc by leadin a trigonal pyramidal tris-thiolate environment. Lead binds about 500 timesmore tightly than zinc to this ligand, but lead is significantly less Lewis acidicthan zinc in this environment.

23. Klotz IM, Urquhart JM, Fiess HA: Interactions of metal ions with thesulfhydryl group of serum albumin. J Am Chem Soc 1952, 74:5537.

24. Mehra RK, Kodati VR, Abdullah R: Chain length-dependent Pb(II)-coordination in phytochelatins. Biochem Biophys Res Commun1995, 215:730-736.

25. Payne JC, ter Horst MA, Godwin HA: Lead fingers: Pb(II) binding to•• structural zinc-binding domains determined directly by monitoring

lead-thiolate charge-transfer bands. J Am Chem Soc 1999,121:6850-6855.

The first direct and quantitative determination of the thermodynamics oflead–protein interactions, using lead–thiolate charge-tranfer bands. Leadbinds tightly to cysteine-rich sites in structural zinc-binding domains and dis-rupts the structure of the proteins.

26. Quintanilla-Vega BH, Hoover DJ, Bal W, Silbergeld EK, Waalkes MP,• Anderson LD: Lead interaction with human protamine (HP2) as a

mechanism of male reproductive toxicity. Chem Res Toxicol 2000,13:594-6000.

UV-vis and circular dichroism spectrocopies are used to investigate theinteractions between lead and HP2. Lead binds to HP2 with a similar affini-ty to that of zinc and interferes with HP2–DNA interactions.

27 Shimoni-Livny L, Glusker JP, Bock CW: Lone pair functionality indivalent lead compounds. Inorg Chem 1998, 37:1853-1867.

28. Hanas JS, Rodgers JS, Bantle JA, Cheng YG: Lead inhibition of• DNA-binding mechanism of Cys(2)His(2) zinc finger proteins. Mol

Pharmacol 1999, 56:982-988.Lead inhibits the interactions between 5S RNA and proteins containingCys2His2 zinc fingers (TFIIIA and SP1). Transcription factors that do notcontain zinc fingers (AP1) are not affected by lead.

29. Kim KA, Chakraborti T, Goldstein GW, Bressler JP: Immediate early• gene expression in PC12 cells exposed to lead: requirement for

protein kinase C. J Neurochem 2000, 74:1140-1146.Lead induces expression of early response genes (c-fos, c-jun, and erg-1) inPC12 cells in a PKC-dependent fashion.

30. Finkelstein Y, Markowitz ME, Rosen JF: Low-level lead-inducedneurotoxicity in children: an update on central nervous systemeffects. Brain Res Brain Res Revs 1998, 27:168-176.

31. Manalis RS, Cooper GP: Presynaptic and postsynaptic effects oflead at the frog neuromuscular junction. Nature 1973, 243:354-356.

32. Habermann E, Crowell K, Janicki P: Lead and other metals cansubstitute for Ca2+ in calmodulin. Arch Toxicol 1983, 54:61-70.

33. Newton AC: Protein kinase C: structure, function, and regulation.J Biol Chem 1995, 270:28495-28498.

34. Rizo J, Sudhof TC: C2-domains, structure and function of a universalCa2+-binding domain. J Biol Chem 1998, 273:15879-15882.

35. Sutton RB, Davletov BA, Berghuis AM, Sudhof TC, Sprang SR:Structure of the first C2 domain of synaptotagmin I: a novel Ca2+/phospholipid-binding fold. Cell 1995, 80:929-938.

36. Sutton RB, Ernst JA, Brunger AT: Crystal structure of the cytosolicC2A-C2B domains of synaptotagmin III. Implications for Ca(+2)-independent snare complex interaction. J Cell Biol 1999,147:589-598.

37. Bouton CMLS, Frelin LP, Forde CE, Godwin HA, Pevsner J:•• Synaptotagmin is a molecular target for lead. J Neurochem 2001,

in press.Lead binds tightly to the calcium site(s) in synaptotagmin but prevents bind-ing of synaptotagmin to its protein partner syntaxin.

38. Brown PO, Botstein D: Exploring the new world of the genomewith DNA microarrays. Nat Genet 1999, 21:33-37.

39. Pandley A, Mann M: Proteomics to study genes and genomics.Nature 2000, 405:837-846.

40. Fleming MD, Trenor CC III, Su MA, Foernzler D, Beier DR,Dietrich WF, Andrews NC: Microcytic anaemia mice have amutation in Nramp2, a candidate iron transporter gene. Nat Genet1997, 16:383-386.

41. Gunshin H, Mackenzie B, Berger UV, Gunshin Y, Romero MF,Boron WF, Nussberger S, Gollan JL, Hediger MA: Cloning andcharacterization of a mammalian proton-coupled metal-iontransporter. Nature 1997, 388:482-488.

42. Tandy S, Williams M, Leggett A, Lopez-Jimenez M, Dedes M,Ramesh B, Srai SK, Sharp P: Nramp2 expression is associatedwith pH-dependent iron uptake across the apical membrane ofhuman intestinal Caco-2 cells. J Biol Chem 2000, 275:1023-1029.

43. Donovan A, Brownlie A, Zhou Y, Shepard J, Pratt SJ, Moynihan J,Paw BH, Drejer A, Barut B, Zapata A et al.: Positional cloning ofzebrafish ferroportin1 identifies a conserved vertebrate ironexporter. Nature 2000, 403:776-781.

44. Rensing C, Sun Y, Mitra B, Rosen BP: Pb(II)-translocating P-typeATPases. J Biol Chem 1998, 273:32614-32617.

45. Sharma R, Rensing C, Rosen BP, Mitra B: The ATP hydrolytic activity• of purified ZntA, a Pb(II)/Cd(II)/Zn(II)-translocating ATPase from

Escherichia coli. J Biol Chem 2000, 275:3873-3878.The zinc transporter ZntA effluxes Pb2+ and confers resistance to lead in E. coli.

46. Erdahl WL, Chapman CJ, Taylor RW, Pfeiffer DR: Ionomycin, a• carboxylic acid ionophore, transports Pb2+ with high selectivity.

J Biol Chem 2000, 275:7071-7079.The ionophore ionomycin selectively transports lead into cells and holdspromise as a useful tool for studying and treating lead poisoning.

47. Petering DH, Huang M, Moteki S, Shaw CF III: Cadmium and leadinteractions with transcription factor IIIA from Xenopus laevis: amodel for zinc finger protein reactions with toxic metal ions andmetallothionein. Mar Environ Res 2000, 50:89-92.

The biological chemistry of lead Godwin 227