Chemotaxis Receptors: A Progress Report on Structure and Function

19
Chemotaxis Receptors: A Progress Report on Structure and Function Sherry L. Mowbray* ,1 and Mats O. J. Sandgren² Department of Molecular Biology, *Swedish Agricultural University, and ²Uppsala University, Box 590, Uppsala Biomedical Centre, S-751 24 Uppsala, Sweden Received July 17, 1998, and in revised form September 30, 1998 Recent biochemical and structural studies have provided many new insights into the structure and function of bacterial chemoreceptors. Aspects of their ligand binding, conformational changes, and interactions with other members of the signaling pathway are being defined at the structural level. It is anticipated that the combined effort will soon provide a detailed, unified view of an entire re- sponse system. r 1998 Academic Press Key Words: chemotaxis; receptor; periplasmic bind- ing proteins; bacterial proteins. INTRODUCTION Chemotaxis, the movement of cells in response to chemical attractants and repellents, is an important survival strategy for many bacterial species and a frequent feature in bacterial pathogenesis. The pro- cess requires the cooperative function of proteins that receive chemical signals and use the informa- tion to direct appropriate swimming behavior. The present review will concentrate on the gram-nega- tive organisms Escherichia coli and Salmonella typhi- murium, because the most is known for these sys- tems, and because the principles are applicable to most others. Excellent general reviews include Blair (1995), Stock and Mowbray (1995), Stock and Surette (1996), and Falke et al. (1997). Particular emphasis will be focused here on the receptors and their relationships to other components of the signaling system. Flagella can rotate either clockwise or counterclock- wise; the former generates nonproductive tumbles, and the latter, smooth swimming. Control of the direction of flagellar rotation turns simple motility into chemotaxis. In the absence of stimuli, a series of smooth swims separated by short tumbles take a bacterium on a random walk through its surround- ings. When cells detect an increasing concentration of an attractant (or decreasing concentration of repellent), they are more likely to swim smoothly. Gradients in the opposite direction cause a greater frequency of tumbling. Thus the bacteria spend more time moving in directions they perceive as favorable. The proteins required for chemotaxis in E. coli are shown in Fig. 1. Initial recognition of an effector molecule requires its binding to a specific receptor on the outside of the cell. Some attractants bind first to a soluble primary receptor, or binding protein, re- cruited from the ABC transport system (Higgins, 1992; Tam and Saier, 1993; Boos and Lucht, 1996). These ligand-loaded binding proteins can interact with, and activate, a particular membrane-bound receptor. Other attractants and repellents are able to bind directly to specific membrane receptors. In either case, it is the membrane-bound protein that actually carries the signal to the cytoplasm and conveys the proper instructions to the rest of the system. All membrane receptors trigger the same intracellular events. Signal processing and output involve a phosphore- lay system. The activity of CheA, a cytoplasmic histidine kinase, is modulated by its interactions with the membrane receptor. CheA is inhibited when an attractant is bound to the receptor and stimu- lated when a repellent signal is given. When acti- vated, CheA can autophosphorylate (using ATP) and then pass the phosphate groups to either of two aspartate autokinases involved in signal output (CheY) and adaptation (CheB). CheW couples CheA to chemoreceptor control by physically linking the proteins in a complex. CheY, the tumble or response regulator, catalyzes its own acceptance of a phos- phate group from CheA on an aspartic acid residue. The default state of a flagellum is smooth swimming in the absence of bound CheY. Only phosphorylated CheY interacts with a flagellum and generates a tumble. CheZ controls the level of phosphorylated CheY by accelerating its rate of autodephosphoryla- tion; thus tumbles rapidly cease when CheA is turned off. Adaptation resets the system by changing the 1 To whom correspondence should be addressed. Fax: (46)-18– 536971. E-mail: [email protected]. JOURNAL OF STRUCTURAL BIOLOGY 124, 257–275 (1998) ARTICLE NO. SB984043 257 1047-8477/98 $25.00 Copyright r 1998 by Academic Press All rights of reproduction in any form reserved.

Transcript of Chemotaxis Receptors: A Progress Report on Structure and Function

Chemotaxis Receptors: A Progress Report on Structure and Function

Sherry L. Mowbray*,1 and Mats O. J. Sandgren†

Department of Molecular Biology, *Swedish Agricultural University, and †Uppsala University, Box 590,Uppsala Biomedical Centre, S-751 24 Uppsala, Sweden

Received July 17, 1998, and in revised form September 30, 1998

Recent biochemical and structural studies haveprovided many new insights into the structure andfunction of bacterial chemoreceptors. Aspects oftheir ligand binding, conformational changes, andinteractions with other members of the signalingpathway are being defined at the structural level. Itis anticipated that the combined effort will soonprovide a detailed, unified view of an entire re-sponse system. r 1998 Academic Press

Key Words: chemotaxis; receptor; periplasmic bind-ing proteins; bacterial proteins.

INTRODUCTION

Chemotaxis, the movement of cells in response tochemical attractants and repellents, is an importantsurvival strategy for many bacterial species and afrequent feature in bacterial pathogenesis. The pro-cess requires the cooperative function of proteinsthat receive chemical signals and use the informa-tion to direct appropriate swimming behavior. Thepresent review will concentrate on the gram-nega-tive organisms Escherichia coli and Salmonella typhi-murium, because the most is known for these sys-tems, and because the principles are applicable tomost others. Excellent general reviews include Blair(1995), Stock and Mowbray (1995), Stock and Surette(1996), and Falke et al. (1997). Particular emphasiswill be focused here on the receptors and theirrelationships to other components of the signalingsystem.

Flagella can rotate either clockwise or counterclock-wise; the former generates nonproductive tumbles,and the latter, smooth swimming. Control of thedirection of flagellar rotation turns simple motilityinto chemotaxis. In the absence of stimuli, a series ofsmooth swims separated by short tumbles take abacterium on a random walk through its surround-ings. When cells detect an increasing concentration

of an attractant (or decreasing concentration ofrepellent), they are more likely to swim smoothly.Gradients in the opposite direction cause a greaterfrequency of tumbling. Thus the bacteria spend moretime moving in directions they perceive as favorable.

The proteins required for chemotaxis in E. coli areshown in Fig. 1. Initial recognition of an effectormolecule requires its binding to a specific receptor onthe outside of the cell. Some attractants bind first toa soluble primary receptor, or binding protein, re-cruited from the ABC transport system (Higgins,1992; Tam and Saier, 1993; Boos and Lucht, 1996).These ligand-loaded binding proteins can interactwith, and activate, a particular membrane-boundreceptor. Other attractants and repellents are able tobind directly to specific membrane receptors. Ineither case, it is the membrane-bound protein thatactually carries the signal to the cytoplasm andconveys the proper instructions to the rest of thesystem. All membrane receptors trigger the sameintracellular events.

Signal processing and output involve a phosphore-lay system. The activity of CheA, a cytoplasmichistidine kinase, is modulated by its interactionswith the membrane receptor. CheA is inhibited whenan attractant is bound to the receptor and stimu-lated when a repellent signal is given. When acti-vated, CheA can autophosphorylate (using ATP) andthen pass the phosphate groups to either of twoaspartate autokinases involved in signal output(CheY) and adaptation (CheB). CheW couples CheAto chemoreceptor control by physically linking theproteins in a complex. CheY, the tumble or responseregulator, catalyzes its own acceptance of a phos-phate group from CheA on an aspartic acid residue.The default state of a flagellum is smooth swimmingin the absence of bound CheY. Only phosphorylatedCheY interacts with a flagellum and generates atumble. CheZ controls the level of phosphorylatedCheY by accelerating its rate of autodephosphoryla-tion; thus tumbles rapidly cease when CheA isturned off.

Adaptation resets the system by changing the1 To whom correspondence should be addressed. Fax: (46)-18–

536971. E-mail: [email protected].

JOURNAL OF STRUCTURAL BIOLOGY 124, 257–275 (1998)ARTICLE NO. SB984043

257 1047-8477/98 $25.00Copyright r 1998 by Academic Press

All rights of reproduction in any form reserved.

signaling (kinase modulating) properties of the recep-tor through reversible methylation. Methylated re-ceptor is a more efficient activator of the kinase. Twoproteins work in tandem to determine receptor meth-ylation under particular conditions. CheR, a methyl-transferase, continuously transfers methyl groupsfrom S-adenosylmethionine molecules to specific siteson the membrane receptor. CheB, the methylester-ase, specifically removes methyl groups from thereceptor. CheB becomes activated by accepting aphosphate group from CheA; this control largelydetermines the observed level of methylation at anytime. The cooperative function of these proteins inchemotaxis is illustrated in Fig. 1.

Chemotaxis includes the first-discovered and best-characterized two-component regulatory system.Similar transmitter (histidine kinase, CheA-like)and receiver (aspartate kinase, CheY-like) modulesare found with many different signaling roles inprokaryotes and eukaryotes (Stock et al., 1995; Falkeet al., 1997; Wurgler-Murphy and Saito, 1997). Thesesystems are indeed so closely related that cross-talkwith the chemotaxis proteins can be demonstrated insome cases, as with the nitrogen regulation system(Ntr; Ninfa et al., 1988).

PERIPLASMIC RECEPTOR STRUCTURE

Only some of the soluble receptors from the ABCtransport systems can participate in chemotaxis.The chemoreceptors of E. coli include the periplas-mic binding proteins for the attractants ribose (RBP),glucose/galactose (GBP), maltose/maltodextrins(MBP), and dipeptides (DBP); a recent report impli-cates the nickel-binding protein (NiBP) in repellentresponses through Tar (De Pina et al., 1995). Theirinteractions with particular transmembrane recep-tors are summarized in Fig. 2. The soluble receptorsare inducible, with final concentrations of 1 mM ormore in the periplasm (Manson et al., 1985). MBPhas been shown to be concentrated near the poles ofthe cell after induction (Dietzel et al., 1978).

A great deal is known about the structures of theperiplasmic binding proteins (for a recent review, seeQuiocho and Ledvina, 1996). Three general classeshave been characterized, with each including one ormore chemoreceptors. RBP and GBP are similar instructure, while MBP (Fig. 3) has a different foldingtopology; DBP and NiBP resemble MBP with anextra domain introduced at the N-terminus of each.Only one small, approximately 60-residue segment

FIG. 1. A response cycle. In the absence of any stimulus, receptors (MCPs) activate the CheA kinase slightly, so that some tumblesoccur and some methylated MCP is present. Binding of attractant (either directly or via a binding protein, BP) causes the MCP to inhibitCheA; the resulting reduction in the level of phosphorylated CheY (assisted by CheZ) allows the flagella to return to their defaultsmooth-swimming state. CheB also loses its phosphate group, reducing methylesterase activity. MCPs eventually become more highlymethylated (by CheR), which allows them to regain their kinase-stimulating activity. The slow time scale of methylation permits smoothswimming to persist for a short time. Removal of the attractant results in MCP that efficiently stimulates CheA kinase activity, and morefrequent tumbles occur. The CheB esterase eventually reduces the methylation of the receptor to the original, prestimulus state. Repellentswork in the opposite manner. The system is extremely sensitive; the binding of a single chemical effector at the surface of the cell can alterof the behavior of the entire cell (Segall et al., 1986).

258 MOWBRAY AND SANDGREN

(corresponding to the N-terminal residues of MBP) issignificantly conserved in sequence and structureamong the three classes; this portion of the periplas-mic receptors has been implicated in transport ratherthan chemotaxis. Despite differences in folding, allhave two globular domains connected by a flexible,multistranded hinge; the structure of MBP illus-trates this in Fig. 3. Monomers are observed insolution studies, even at high protein concentrations(Shilton et al., 1996).

Conformational changes of the binding proteinsare essential for their function. Structures of RBP(Mowbray and Cole, 1992; Bjorkman and Mowbray,1998), MBP (Spurlino et al., 1991; Sharff et al., 1992)and DBP (Dunten and Mowbray, 1995; Nickitenko etal., 1995) have been solved in both closed, ligand-bound and open, ligand-free forms. The two statesdiffer by a large, almost rigid-body rotation of thedomains relative to each other. Comparison of theknown structures of open and closed binding pro-

FIG. 2. Periplasmic and membrane chemoreceptors of E. coli. The membrane-bound MCPs, Tar (aspartate receptor), Tsr (serinereceptor), Trg (ribose/glucose/galactose receptor),and Tap (dipeptide receptor), can interact with some ligands directly, and other responsesrequire the periplasmic binding proteins, MBP (maltose-binding protein), GBP, (glucose/galactose-binding protein), RBP (ribose-bindingprotein), or DBP (dipeptide-binding protein). The repellent response to nickel has been reported to work through NiBP (nickel-bindingprotein) and Tar.

259CHEMOTAXIS RECEPTORS

teins using the small conserved portion has shownthat each of the opening motions involves a similarrotation axis passing near the hinge (Flocco andMowbray, 1998). Each open form is stabilized bycontacts between the two domains on the outside ofthe protein, just as closed forms are stabilized byinteractions of each domain with the ligand (Floccoand Mowbray, 1998).

The primary function of these large conforma-tional changes is to provide distinct species thathave distinguishable interactions with chemotaxisand transport partners. The population of forms insolution, however, includes closed ligand-free recep-tors (Flocco and Mowbray, 1994), open ligand-boundreceptors (Sack et al., 1989), and probably manyintermediate (Bjorkman and Mowbray, 1998) and

FIG. 3. Conformational changes of MBP. Open, ligand-free (Sharff et al., 1992) and closed, ligand-bound (Spurlino et al., 1991) forms ofMBP (N-terminal domain at right) with the regions implicated in chemotaxis marked in gray.

260 MOWBRAY AND SANDGREN

less-defined species (Butler and Falke, 1996; Care-aga and Falke, 1992). Ligand should probably beviewed as increasing the population of similar closedforms, rather than causing one all-or-nothing confor-mational change. The exact nature of the open formis probably unimportant, as its function is mostly toreduce the competition of ‘‘inappropriate’’ forms forthe attention of membrane components of chemo-taxis and transport.

MEMBRANE RECEPTOR STRUCTURE

Most transmembrane signaling is carried out by aset of homologous proteins often called MCPs2

(methyl-accepting chemotaxis proteins) because oftheir adaptational methylation. Four closely relatedMCPs are found in E. coli. The responses they directand their specific interactions with soluble receptorsare summarized in Fig. 2. Salmonella has a slightlydifferent set, with Tap replaced by Tcp, a citrate/phenol sensor (Yamamoto and Imae, 1993).

Tar, the Aspartate Receptor

The aspartate receptor, Tar, serves as the primaryexample here. This 60-kDa protein can generatesignals in response to direct binding of the at-tractants aspartate and glutamate (Clarke and Kosh-land, 1979). Although Tar from E. coli and Salmo-nella have 79% sequence identity, only E. coli Tarcan respond to MBP-mal as an attractant (Dahl andManson, 1985; Mizuno et al., 1986) and the ions Ni21

and Co21 as repellents (Reader et al., 1979). A role inthermosensing is shared with the serine receptor,Tsr (Mizuno and Imae, 1984). There are about 2500copies of Tar per cell (Clarke and Koshland, 1979).The absence of cysteine residues in wild-type Tar hasenabled the use of cysteine mutants to test functionand proximity of particular residues.

Homodimers

That dimers are the smallest functional unit of Tarstructure was first established by disulfide cross-linking (with cysteines introduced by mutation) andsedimentation measurements. Neither the bindingof aspartate and nor the methylation altered theobserved size, although aspartate did prevent ex-change of subunits between dimers (Milligan andKoshland, 1988). Tetramers had been previouslyreported as a major species in the membrane usingmore promiscuous cross-linking techniques (Chelskyand Dahlquist, 1980) and may represent a validmechanism for subunit exchange in the absence ofligand. There is no evidence that heterodimers of Tar

can form with other MCPs; even heterodimers ofSalmonella and E. coli Tar are not observed (Milli-gan and Koshland, 1988).

Basic Structure

That the MCPs are remarkably flexible was sug-gested by disulfide trapping experiments with theperiplasmic region of Tar (Falke and Koshland,1987), which were later extended to the transmem-brane and cytoplasmic regions (Danielson et al.,1997). To a first approximation, any introducedcysteine can eventually cross-link with any other inthe vicinity. This flexibility has made structuralwork with these proteins extremely difficult.

No x-ray structure exists for the intact molecule,and electron microscopic studies have met withlimited success. Two-dimensional crystals of Trg(Barnakov et al., 1994) exhibited a P4 symmetry,which is difficult to consolidate with a dimeric struc-ture. It was proposed that the two-dimensionalcrystals might represent an antiparallel arrange-ment of two symmetrical dimers that shows appar-ent fourfold symmetry in the averaged projection.The 150-A spacing in multilamellar stacks indicatesthat the maximum length for any single solubleportion of the receptor is 120 A (assuming a30-A-thick membrane).

Early circular dichroic measurements showed thatthe helical content of Tar is very high (,80%; Fosteret al., 1985), a theme that has persisted in all laterstudies. Most available structural information hasresulted from a determined divide-and-conquer strat-egy utilizing a variety of methods. The compositeview of the Tar dimer in Fig. 4 thus represents thesynthesis of many results. In the descriptions of theindividual regions, a prime (8) notation will be usedto designate the second subunit of a dimer.

N-terminus. The N-terminal methionine of Taris blocked with a formyl group (Milligan and Kosh-land, 1990). The small N-terminal cytoplasmic seg-ment of Tar (residues 1–6) is believed to be helical,with residues 4 and 48 closest together near thedimer interface (Stoddard et al., 1992). Although the4/48 distance is reduced slightly on binding aspartate(Stoddard et al., 1992), this portion of the receptorcan be altered greatly without major functionaleffects (Chen and Koshland, 1995).

Periplasmic region. The role of the periplasmicregion in ligand-binding (Tar residues 31–188) isreflected in the relatively low level of sequenceconservation in this part of the MCPs; bindingdifferent chemoeffectors requires variations in se-quence, although the same basic structure is main-tained.

This portion of the Tar structure is well character-ized. An 18-kDa soluble fragment of the Salmonella

2 Abbreviations used: MCP, methyl-accepting chemotaxis pro-tein; TM1, transmembrane segment 1; TM2, transmembranesegment 2.

261CHEMOTAXIS RECEPTORS

262

protein was produced and shown to bind aspartatewith high affinity (Milligan and Koshland, 1993).Although monomers predominated at low proteinconcentrations in the absence of aspartate, highprotein concentrations or the presence of aspartatefavored the formation of dimers. X-ray structures ofperiplasmic fragments of Salmonella and E. colihave been solved (Milburn et al., 1991; Bowie et al.,1995; Yeh et al., 1996). This fragment of Tar is asymmetric dimer in the absence of ligand. Eachsubunit is composed of an antiparallel four-helixbundle roughly 20 A in diameter, with a1 and a4making extensive contacts throughout their 70-Alength (Milburn et al., 1991; Yeh et al., 1993; Bowie etal., 1995). NMR studies of a periplasmic piece of Tarhave confirmed that the x-ray structure is a goodmodel for the MCP in solution (Danielson et al.,1994).

These structures have provided the basis for mod-eling the structure of Tsr (Jeffery and Koshland,1993). Studies of intact, membrane-bound Tsr bysolid-state NMR validate its use (Wang et al., 1997).

Transmembrane segments. Two segments of Tarcross the membrane: residues 7–30 (TM1; immedi-ately before a1 of the periplasmic region) and resi-dues 189–212 (TM2; directly following a4). By ex-trapolation from the structure of the periplasmicdomain, the membrane segments of the dimer wereproposed to form a pseudo four-helix bundle (Mil-burn et al., 1991). This model is consistent with theresults of cross-linking studies with introduced cyste-ine residues (Lynch and Koshland, 1991; Pakula andSimon, 1992; Chervitz et al., 1995; Chervitz andFalke, 1995). Both transmembrane segments show aclear helical pattern. TM1s provide the closest asso-ciation on the dimer interface, and clear contactsexist between TM1 and TM2 within each subunit.Similar data have been obtained for Trg, both in vivoand in vitro (Lee et al., 1994; Lee and Hazelbauer,1995; Hughson et al., 1997).

TM1s and the cytoplasmic residues immediatelyfollowing them are thought to form a continuouscoiled coil with a1/a18 (Scott and Stoddard, 1994).TM2s appear to be normal a-helices (Chervitz andFalke, 1996) that make contact with the TM1s oftheir respective domains, but not with each other.

Cytoplasmic region. The cytoplasmic region isabout 37 kDa in size and includes residues 213–553

of Tar. This is the most conserved part of the MCPs(sequence identity ,70% between Tar and Tsr),reflecting the fact that all receptors function withinthe same context of intracellular signaling/adapta-tion. It is also the part of the receptor that is leastunderstood.

No X-ray or NMR structure has been obtained forany portion of the cytoplasmic region, apparentlybecause it is highly dynamic (Seeley et al., 1996). Thebroad picture available at present has therefore beenderived from a number of different types of studies.Circular dichroic measurements again showed alarge helical component (Mowbray et al., 1985).Previous success in predicting the general arrange-ment of the periplasmic region (Moe and Koshland,1986) led to a similar attempt for the cytoplasmicpart of the receptors. Sequence comparisons com-bined with secondary structure prediction producedthe helical assignments that are shown in Fig. 4 (LeMoual and Koshland, 1996; Danielson et al., 1997)and used as navigational aids in the following discus-sion. These helices are illustrated in proportion totheir predicted length (on the same scale as theperiplasmic fragment) and in the ‘‘dimeric’’ arrange-ment suggested by experimental data, although noparticular model is implied for the disposition of a7and a8. The available data on size (Long and Weis,1992a; Barnakov et al., 1994; Liu et al., 1997) hintthat a7 and a8 are not fully extended and may foldback in some way onto a6 and a9.

Residues corresponding to 213–259 of Tar arecommonly referred to as the linker. This region issomewhat conserved (Le Moual and Koshland, 1996),and its integrity is critical to receptor function(Danielson et al., 1997; Ames et al., 1988; Ames andParkinson, 1988; Gardina and Manson, 1996). Cyste-ine and disulfide scanning (Butler and Falke, 1998)indicated that it includes two amphiphilic helices,one of which is an extension of a4, while the other(a5) is at the dimer interface. Helix a4 includes aconserved proline that may encourage a kink. Theregion between a4 and a5 (229–242) is apparentlymore ordered and less exposed than either of thenearby helices; other studies indicate that this is ahot spot for mutations that disrupt signaling (Ameset al., 1988).

A protease-accessible site at residue 259 marksthe end of a5 (Mowbray et al., 1985). Although such

FIG. 4. Composite structure of a Tar dimer. Various experimental information was combined with the x-ray structure of theaspartate-bound periplasmic region (Brookhaven Protein Data Bank entry 1WAT) to give an overall schematic. The helical regionsobserved in the x-ray structure, combined with those suggested for TM1 and TM2, are a1 (4–76), a2 (86–109), a3 (116–144), and a4(146–212). A single bound aspartate is shown as a space-filling model. The additional helical regions proposed by Le Moual and Koshland(1996) and Danielson et al. (1997) are shown at approximately the same scale as the periplasmic part, assuming a rise/residue of 1.5 A: a4(212–228), a5 (247–261), a6 (265–337), a7 (345–387), a8 (400–426), and a9 (438–516). Helices a6 and a68 are shown coiled around eachother with a pitch of ,140 A. Methylation sites at residues 295, 302, 309, and 491 are indicated by stars.

263CHEMOTAXIS RECEPTORS

protease cleavage is abolished by conditions thatfavor helix formation (such as high concentrations ofglycerol), analysis of sequences (Le Moual and Kosh-land, 1996; Danielson et al., 1997) suggests that a5 isdistinct from the following helix, a6. The division ofTar into two stable structural units with limitedproteolysis led to studies utilizing a convenientnearby restriction site to prepare C-terminal frag-ments consisting of residues 258–553 (Kaplan andSimon, 1988). This portion of the receptor includestwo regions involved in methylation, linked by ahighly conserved region implicated in interactionswith CheA and CheW; it terminates in a distinctC-terminal ‘‘tail.’’ Such fragments appear to be highlyhelical in CD measurements (Mowbray et al., 1985;Wu et al., 1995) and move as much larger species ongel filtration chromatography (Mowbray et al., 1985;Kaplan and Simon, 1988; Long and Weis, 1992a,b). Amonomer–dimer equilibrium is observed, with mono-mers having an apparent Stokes radius of about 30A, and dimers, of about 40 A; the relative proportionsof monomer and dimer are greatly altered by solu-tion conditions as well as by mutation (Long andWeis, 1992a). Dissociation of the dimers seems torequire some unfolding (Wu et al., 1995), as wouldnot be surprising if they intertwine in the mannershown in Fig. 4. Structural work with these frag-ments has been hampered by their extreme flexibil-ity. They are highly sensitive to protease (Mowbrayet al., 1985), even in dimeric form (Seeley et al.,1996), and unusually mobile according to severalspectroscopic criteria (Seeley et al., 1996). The iso-lated fragments crystallize, but the crystals havepoor diffraction characteristics (M. O. J. Sandgren, B. H.Shilton, and S. L. Mowbray, unpublished results).

Reversible methylation of MCPs occurs at four tosix glutamate residues (Springer and Koshland,1977; Terwilliger and Koshland, 1984; Nowlin et al.,1987; Rice and Dahlquist, 1991), corresponding toGln295, Glu302, Gln309, and Glu491 in Tar. Thesites originally coded for as glutamine residues aremade available for methylation by a secondary deami-dating activity of the chemotaxis methylesterase(Rollins and Dahlquist, 1981; Sherris and Parkin-son, 1981). The methylation regions were earlyproposed to be helical, based on sequence patternsnear the modification sites (Kehry and Dahlquist,1982; Terwilliger et al., 1983), a model that issupported by sequence comparisons and disulfidescanning (Danielson et al., 1997). The methylationregions are currently thought to be part of the longerhelices a6 and a9. Patterns consistent with theformation of coiled coils are observed for MCP resi-dues from approximately 290 to 330 and 475 to 520(Lupas et al., 1991). Helices a6 and a9 are suggestedto be antiparallel to each other, forming a four-helix

bundle that includes the corresponding helices of theother subunit (Stock et al., 1991; Surette and Stock,1996). Helices a6 and a68 have been shown to make aparallel contact (Danielson et al., 1997), while a9 anda98 have yet to be examined in detail. An intra-subunit coiled coil should be stabilized by free gluta-mates (Lupas et al., 1991), while the bundle may befavored by neutralization of these charges withmethyl or amide groups (Stock et al., 1991). Such anarrangement would be expected to ‘‘wrap around’’itself with a left-handed pitch of about 140 A (Fig. 4).

Various fragments containing a highly conserved‘‘core’’ (approximately residues 350–425, that is, a7and a8 and the loop connecting them) have beencalled the ‘‘signaling domain’’ with reference to rolesin binding CheA and CheW as well as kinase modula-tion. An independently folding fragment of Tsr (resi-dues 350–470) has been shown to stimulate CheA ina CheW-dependent manner (Ames et al., 1996);combined with similar observations for an overlap-ping section of Tar (Surette and Stock, 1996), itappears that residues from ,350 to 455 actuallyform the minimum folding unit.

A model where a7 and a8 are extended in directlines from a6 and a9, respectively, seems incompat-ible with the available data on the length of thereceptor, and so more compact arrangements shouldbe considered. A coil pattern analysis (see, e.g., Stocket al., 1991) shows similar amphiphilic patterns inthis region of the sequence and in the periplasmicdomain, suggesting the possibility a four-helixbundle, with individual helices in the regions of340–365 (start of a7), 365–385 (end of a7), 390–430(a8), and 430–470 (start of a9). The beginning and/orending helices could be continuous with a6/a9. An-other possibility involves the simple folding up of a7and a8 onto a6 and a9. An intrasubunit bundle ofthis type would merge into a dimer bundle in the‘‘upper’’ reaches of a6 and a9, since a7 and a8 areshorter (Fig. 4 illustrates this possibility); again, thesimilarity between this arrangement and that ofthe periplasmic domain is clear. Such folding mightbe affected by methylation, since the modificationsites could be buried in the folded form. It is quitepossible that more than one type of helical arrange-ment will ultimately be found; the accessibility ofmultiple states seems to be a common, and impor-tant, feature of proteins containing coiled coils(Lupas, 1996).

A short region of indeterminate structure (resi-dues 515–553) follows a9. This C-terminal ‘‘tail’’ iseasily removed with protease (M. O. J. Sandgren,unpublished results) and appears highly mobile inspectroscopic measurements (Seeley et al., 1996).The last five residues of Tar have now been shown to

264 MOWBRAY AND SANDGREN

be essential for methyltransferase tethering (seebelow).

Other receptors. Similar MCPs are found inmany bacterial species (Dahl and Manson, 1985;Alam and Hazelbauer, 1991), including nonmotile(Dahl et al., 1989) as well as gliding bacteria (Mor-gan et al., 1993). The general topology places theseMCPs in the type I family of transmembrane recep-tors (Heldin, 1995), a relationship that has beensupported by the construction of functional chimericreceptors of Tar with the insulin receptor (Ellis et al.,1986; Moe et al., 1989; Biemann et al., 1996). To-gether with the discovery that a dimer of CheA is theactive form (Surette et al., 1996), this idea hasprovoked much discussion of transmembrane signal-ing by monomer–dimer mechanisms, as is appar-ently the case for many eukaryotic type I receptors.However, it appears that as for the insulin receptor,allosteric mechanisms predominate for the MCPs.

A fifth, somewhat different, MCP of E. coli hasrecently been characterized (Rebbapragada et al.,1997; Zhulin et al., 1997; Bibikov et al., 1997; Grisha-nin and Bibikov, 1997). The Aer receptor works insynergy with Tsr in aerotaxis and energy sensing.Aer has a cytoplasmic signal-transducing regionsimilar to the more classical MCPs, but a cytoplas-mic sensing region of a completely distinct design(containing a noncovalently associated FAD). SomeMCPs in other organisms have no periplasmic por-tions and either one or two membrane-spanningsegments (Krah et al., 1994; Spudich, 1994; LeMoual and Koshland, 1996), while others are solubletransducers including only the cytoplasmic signalingand methylation regions (Zhang et al., 1996; Broounet al., 1997). It is presumed that the cytoplasmicportions of all of these signaling systems will followthe same general model.

The phosphotransferase system also uses CheAand CheY to send intracellular signals, effectivelyusing enzymes I and II as receptors (Lux et al., 1995).

Receptors Are Concentrated at Poles of Cells

Associations of large numbers of receptor dimersare likely to play an important role in their function.About 80% of the Tar in wild-type E. coli is foundclustered in large patches at the poles of the cell.Formation of the patches has been shown to bedependent on the presence of CheW and, to a lesserextent, CheA; CheY and CheZ are apparently notnecessary (Maddock and Shapiro, 1993). Althoughpolar clustering requires the intracellular (C-termi-nal) portions of the MCP (Alley et al., 1992, 1993),the origins of the phenomenon are not yet clear.Flagella are not localized in this way, nor are MCPsspecifically clustered near the flagella (Nathan et al.,

1986). A correlation has, however, been noted with asimilar disposition of MBP (Maddock et al., 1993);expression of MBP had been linked earlier withpolar cap formation (Boos and Staehelin, 1981; Diet-zel et al., 1978).

Smaller, regular arrays were observed in electronmicroscopic studies of a cytoplasmic construct of Tarwhere oligomerization was induced with leucinezippers (Liu et al., 1997). This association occurredin the absence of CheW and CheA, a long thin shapeof the cytoplasmic fragments allowing their packinginto compact bundles. Each individual receptor re-gion appeared to account for approximately 140 A inlength, similar to the maximum of 120 A suggestedby electron microscopy of Trg. This would correspondroughly to the length of cytoplasmic portions of a4,a5 and a6 arranged in extended form. A ‘‘nonex-tended’’ arrangement in the cytoplasmic regionsmight be expected to ease packing of intact MCPsinto such arrays, by making the periplasmic andcytoplasmic regions more similar in size and shape.CheA and CheW bound at the ends of the bundles,with approximately 14 receptor dimers being associ-ated with six CheW molecules and two CheA dimers.Clustering was clearly linked to kinase activation,and both kinase activity and clustering were in-creased by higher levels of methylation (Liu et al.,1997). Methylation may act to promote bundle forma-tion by neutralizing the negative charges of themodified residues, as suggested above for dimers.

LIGAND BINDING ALTERS AN MCP

The measured dissociation constant for aspartatebinding to Tar is approximately 1 µM in membrane-bound and solubilized intact receptors, as well asperiplasmic fragments (Aksamit et al., 1975; Clarkeand Koshland, 1979; Foster et al., 1985; Mowbray etal., 1985; Milligan and Koshland, 1993; Biemannand Koshland, 1994; Danielson et al., 1994). Thisrepresents binding to the first of two possible sites onthe dimer; a second aspartate binds with loweraffinity (Milligan and Koshland, 1993; Biemann andKoshland, 1994). The relative affinity of the two sitesfor aspartate is not constant. While binding at thesecond site occurs fairly readily in the SalmonellaTar, the negative cooperativity is more absolute in E.coli Tar (Biemann and Koshland, 1994), as well as inTsr (Lin et al., 1994), where only a single ligandmolecule binds to each dimer.

These results can be understood in terms ofchanges observed in the x-ray structures on aspar-tate binding. The binding site has contributions fromboth subunits (Fig. 5; Milburn et al., 1991). Whenaspartate binds to one site between the two initiallyequivalent subunits, they rearrange to a nonsym-

265CHEMOTAXIS RECEPTORS

metric dimer (Milburn et al., 1991; Yeh et al., 1996).Local alterations at the second site, and more globalchanges within the dimer, dictate that affinity at thesecond site will be different (Scott and Stoddard,1994). Binding of a second aspartate to wild-typeSalmonella Tar does not appear to cause any addi-tional changes in the overall structure compared tothe single-site binding (Yeh et al., 1996), althoughthe form observed could be biased by crystal packing.That this situation is ‘‘adjustable’’ is indicated notonly by the E. coli/Salmonella differences, but by thefact that mutation of Salmonella Tar at Ser68 in thedimer interface can give rise to positive, negative, orno cooperativity (Kolodziej et al., 1996).

A simple monomer–dimer equilibrium cannot ex-plain transmembrane signaling; many cross-linkeddimers can signal (Falke and Koshland, 1987). In-stead, changes within an individual subunit appearto be critical (Scott and Stoddard, 1994; Danielson etal., 1994; Chervitz and Falke, 1996). Main- andside-chain groups on a1 and a4 of one subunitinteract almost exclusively with main-chain groups

of the aspartate ligand. Complementary groups ona18 of the other subunit make most contacts with theligand side chain. When aspartate enters the site, a4moves about 1.6 A toward the membrane relative toa1, and is tilted by about 5°, in a ‘‘swinging piston’’motion (Fig. 5; Chervitz and Falke, 1996). In thisway, a4 moves about 1 A with respect to the helices ofthe other subunit, which itself is unchanged. Thedimer interface, mainly a function of a1/a18 interac-tions, remains relatively static. Inspection of thex-ray structures of periplasmic fragments of E. coliTar with and without bound aspartate (Bowie et al.,1995) suggests a similar situation. NMR studiesdemonstrate further that although aspartate andglutamate generate somewhat different conforma-tions locally at the binding site, they cause a similarintrasubunit conformational signal to be sent throughthe receptor (Danielson et al., 1994). Electron para-magnetic resonance measurements with intact, mem-brane-bound Salmonella Tar also indicate a relativemovement of a4, which may be accompanied by aslightly closer approach of a1 and a18 (Ottemann etal., 1998).

Motion of an a4 is transmitted to the membranesegment (TM2) to which it is attached and thus cancarry a signal across the membrane. Cross-linkingstudies confirm that the interface between TM1/TM18 plays a mainly static, structural role, while themotion of TM2 relative to TM1 is most critical inkinase modulation (Lynch and Koshland, 1991;Danielson et al., 1994; Scott and Stoddard, 1994;Chervitz et al., 1995; Chervitz and Falke, 1995).These changes are similar to the helix-on-helix slid-ing motions observed to affect form and function inother proteins (Chothia and Lesk, 1985). Modeling oflock-off disulfides suggests that they generate thesame membrane-ward motion of a4 as the binding ofaspartate, while lock-on disulfides appear to cause amotion in the opposite direction (Chervitz and Falke,1996).

Such motions can carry binding signals to thecytoplasm, but can information from the cytoplasmbe sent back to the periplasmic regions? Most studies(Springer et al., 1979; Clarke and Koshland, 1979;Dunten and Koshland, 1991; Borkovich et al., 1992;Lin et al., 1994; Iwama et al., 1997) suggest thatmethylation of the receptors has only a small effecton the affinity for ligand, at least at the first site.Differences would also be expected in the ligandaffinity of MCPs with varying kinase-modulationproperties, but have been similarly hard to pinpoint.Although clustering may allow small effects withinany individual receptor to combine synergistically(Liu et al., 1997), the impact of parameters specific tothe in vivo system is very hard to assess.

FIG. 5. The aspartate-binding site of Tar. Arg64 (on a1) formshydrogen bonds with the a-carboxylate of the aspartate ligand,while the a-amino group is bound by Thr154 and a nitrocationhole made up of the backbone carbonyl oxygens of residues 149and 152 (all on a4 or the loop just before it). Arg698 and Arg738 arelocated on a18 of the other subunit and make contact with both a-and g-carboxylates. The position of a4 and the preceding loopprior to aspartate binding is indicated by the lighter backbone.Not shown are the side chain of Tyr149, which makes van derWaals contacts as well as hydrogen bonds to both a- and g-carbox-ylates indirectly through water molecules, and Ser68 whichinteracts indirectly via water. The figure was made using ProteinData Bank entries 1LIH and 2LIG.

266 MOWBRAY AND SANDGREN

INTERACTIONS BETWEEN MEMBRANEAND PERIPLASMIC RECEPTORS

Trg and Tap recognize only binding proteins, andTsr only small molecules, while E. coli Tar is uniquelycapable of recognizing both small molecule andbinding protein ligands (Fig. 2). The available stud-ies indicate that the mechanism of transmembranesignaling is similar for the two types of attractant(Lee et al., 1994, 1995a,b; Lee and Hazelbauer, 1995;Hughson and Hazelbauer, 1996; Baumgartner andHazelbauer, 1996; Hughson et al., 1997; Gardina etal., 1997). Mutants of Trg can be found, for example,that continuously send kinase-modulating signals,as well as ones that are insensitive to bindingprotein (Yaghmai and Hazelbauer, 1992). Bindingprotein-dependent responses are inherently weaker(e.g., Mowbray and Koshland, 1987) even when theMCP is apparently saturated (Manson et al., 1985),implying that binding protein-induced conforma-tional changes are either smaller or otherwise lesseffective. Since the primary effect of attractant bind-ing is to inactivate the kinase, there may indeed bemultiple ways to accomplish this task and variationsin their efficiency.

The dissociation constant for the MBP–Tar interac-tion been suggested by in vivo expression studies tobe of the order of 250 mM (Manson et al., 1985),although it is not clear how polar localization of MBPwithin the periplasm might affect this estimate. Acombination of mutational studies of MBP and Tar(Manson and Kossman, 1986; Kossman et al., 1988;Zhang et al., 1992; Gardina et al., 1997) and com-puter modeling gave rise to the view of the complexpresented in Fig. 6 (Zhang et al., 1998). As with asimilar earlier model (Stoddard and Koshland, 1992,1993), a single molecule of MBP is proposed tointeract with a dimer of E. coli Tar through the loopsat the ‘‘top’’ of the MCP; the present model, however,agrees better with the effects of known mutations.The surface of MBP that is implicated in chemotaxisis similar in size and disposition relative to the sitesimportant in transport, although not in structure, tothat proposed for the RBP/Trg complex (Binnie et al.,1992). Interactions of MBP with the a3–a4 loop ofTar could easily give rise to conformational changeswithin one subunit of the MCP that are analogous tothose caused by aspartate binding (Fig. 6).

This model also suggests an explanation for theobservation that maltose and aspartate responses ofthe E. coli receptor are quite independent: maltosecan generate a signal even in the presence of saturat-ing aspartate and vice versa (Wolff, 1983; Mowbrayand Koshland, 1987). Previous experimental datasuggested that the two signals are carried in parallelby different subunits of the same receptor dimer(Gardina et al., 1997). If the relationship of a18 and

a48 is unchanged in the second subunit of an singlyaspartate-bound dimer, MBP should be able to usethis unactivated subunit to trigger a response evenwhen one molecule of aspartate is present. Minordifferences in the structures of Salmonella and E.coli Tar would be sufficient to dictate that, when thedimers have bound to a single molecule of aspartate,the former protein would be competent to bind asecond aspartate, while the latter would prefer tointeract with MBP.

MODULATION OF CHEA BY MCPS

Clearly, motions of a single a4 within an MCPdimer can affect both intrasubunit and intersubunitstructure in the cytoplasmic regions, with correspond-ing consequences for receptor interactions with CheAand CheW. That the intrasubunit effects dominatewithin any individual MCP dimer is suggested bystudies showing kinase modulation (Gardina andManson, 1996; Tatsuno et al., 1996) and methylation(Milligan and Koshland, 1991) with receptor dimerswhere one subunit has been truncated (but stillincludes the linker). Since the activation of CheA isbelieved to involve its dimeric form (Surette et al.,1996; Levit et al., 1996), these results are difficult toexplain within the context of a one receptor dimer:one CheA dimer:two CheW monomer complex thathas been demonstrated in vitro (Gegner et al., 1992).They are, however, more easily rationalized for thetype of multi-MCP aggregates that seem to be neces-sary for maximal stimulation of CheA (Liu et al.,1997). (The truncated receptors have not, however,been demonstrated to form clusters.) In reality, theamounts of CheA and CheW present in the cell areinsufficient to form 2:2:2 complexes with all avail-able MCP (data on quantities summarized in Falkeet al., 1997, and Stock et al., 1991), suggesting that,in vivo, a mixture of larger and smaller complexes islikely to exist. Indeed, the spectrum of activationlevels obtainable from the various species may be animportant aspect of receptor function.

In any case, conformational signals associatedwith ligand binding must be converted into changesin CheA activity, presumably via changes in MCPstructure near a7 and a8 (the conserved signalingdomain). The ‘‘crude’’ nature of the signal required atthe periplasmic side is suggested by the success ofmany functional chimeric constructs. Indeed, a recep-tor with both transmembrane segments removed isstill capable of sending a signal (Ottemann andKoshland, 1997).

Observations that CheA by itself is in equilibriumbetween an inactive monomeric form and an activedimer (Surette et al., 1996) suggested that receptor-induced superactivation of CheA (Borkovich et al.,1989; Ninfa et al., 1991; Borkovich and Simon, 1990)

267CHEMOTAXIS RECEPTORS

involves some variation on the dimer theme. Incontrast, the inhibition of CheA by attractant-boundreceptors may involve a return to a monomer-likestate of CheA. The mechanisms by which eitheractivation or inhibition of CheA occur are far fromclear. Control does not appear to rely on association/dissociation of MCP/CheA/CheW complexes, as their

lifetime is too long (Gegner et al., 1992; Schuster etal., 1993). Dimers of Tar with appropriate cross-linksbetween their a1/TM1s can affect the kinase eitherpositively or negatively (Chervitz and Falke, 1995),further suggesting that changes in function areaccomplished allosterically rather than through asso-ciation/dissociation of the cytoplasmic regions. Two

FIG. 6. Model for the MBP/Tar complex. A Tar dimer (light blues) and MBP (gold) are shown with ribbon representations. The view ofTar differs slightly from that used in Fig. 4; the first domain of MBP is at left. Red highlights residues of MBP (41, 45, 46, 49, 50, 53, 55, 342,345, 354, 359, and 367) and Tar (69, 73, 75–77, 83, 143, 145, 147, 149, and 150) that have been implicated in maltose chemotaxis (Kossmanet al., 1988; Zhang et al., 1992; Gardina et al., 1997). The model coordinates used to prepare this figure were kindly provided by Y. Zhang,P. J. Gardina, J. Christopher, and M. D. Manson of Texas A&M.

268 MOWBRAY AND SANDGREN

distinctly different types of ‘‘cytoplasmic’’ dimershave indeed been reported. Dimers in which a6/a68contacts are directed toward a specific coiled-coilpattern are associated with CheA activation (Suretteand Stock, 1996; Cochran and Kim, 1996). Thepresence of a9 and a98 somehow reduces this stimu-lation (Ames and Parkinson, 1994; Surette andStock, 1996), perhaps by changing the relationshipof a6 and a68. Another type of dimer is observed(Long and Weis, 1992a,b) for cytoplasmic constructsthat represent smooth-swimming (CheA-inhibiting)mutants of Tar (Mutoh et al., 1986; Oosawa et al.,1988). This second type of dimer has been shown inat least one case to be dominant when expressedtogether with intact receptor, indicating effectivebinding of CheA and/or CheW (Oosawa et al., 1988).It could, therefore, represent an important alternatestate of the MCP, much like the open forms of thebinding proteins. Fundamentally different interac-tions between CheA and the MCPs in the two stateshave been suggested by other studies, with thekinase-on interaction being more dependent on CheW(Ames and Parkinson, 1994). Such a two-state modelis illustrated schematically in Fig. 7. Although itseems likely that there are a limited number of waysof turning the kinase on, there is almost certainlymore than one way of turning it off. The piston-likemotion seen in the structures of the periplasmicfragments illustrates one way of inactivating thekinase and rotational movements of a6/a68 withrespect to each other represents another possibility(Cochran and Kim, 1996). Kinase inactivation result-ing from small-molecule or binding protein associa-tions may reflect different combinations of these two

components, although rotational inactivation hasnot been demonstrated in intact receptors.

Increased methylation of an MCP acts in theopposite manner to attractant binding, i.e., morehighly methylated MCP stimulates the kinase moreefficiently (Ninfa et al., 1991; Borkovich et al., 1992).If aspartate binding works to weaken the bundling ofa6s and a9s in the dimer, then glutamate modifica-tion may well favor the interaction by reducingelectrostatic repulsion between the helices (Stock etal., 1991). In this situation, it is easy to imaginemultiple methylation events having qualitativelysimilar results.

Dramatic changes in secondary and tertiary struc-ture seem to be rather common for proteins involvingcoiled coils (Lupas, 1996). It has been proposed (Brayet al., 1998) that such changes within a dimer couldbe passed on to ‘‘infect’’ neighbors, making clusteredMCPs supersensitive to the effects of aspartatebinding. A mixture of different, and varying, levels ofaggregation seems most consistent with the charac-teristics of sensing in vivo. Further, it is predictedthat clustering will be extensive at low concentra-tions of aspartate and much less at high concentra-tions of aspartate. Other workers have suggestedthat slight fluctuations in the levels of CheW provideanother means of controlling the formation, and thuseffects, of clusters (Liu et al., 1997).

The behavior of any cell is thus the result of abalance of stresses that originate in the periplasmwith compensating cytoplasmic interactions. A re-sponse, positive or negative, is due to a temporarydisturbance in that balance. The probability of the

FIG. 7. A two-state model. Ligand binding and covalent modification appear to act by altering the probability of CheA kinaseactivation. The interconversion between the two states in shown here (from the ‘‘top’’ of the dimer) as a dramatic (30°) rotation, fordemonstration purposes. The activation of CheA could, of course, be ‘‘interdimeric,’’ but the basic ideas remain the same. Experimentalevidence suggests that alterations in the regions of the a6/a9 contacts, and the associated a6/a68 contacts, are involved. Both kinase-on andkinase-off states of the receptor appear to involve a5/a58 (Butler and Falke, 1998) and a6/a68 contacts (Danielson et al., 1997; Cochran andKim, 1996), but the relationship to a9 and a98 to the rest of the structure is not really established for either case.

269CHEMOTAXIS RECEPTORS

kinase being activated under given conditions is thekey issue.

INTERACTIONS WITH METHYLTRANSFERASEAND METHYLESTERASE

Great progress has been made recently in under-standing the interactions of MCPs with the enzymesthat covalently modify them. Realization that meth-yltransferase actually binds most tightly to a five-residue sequence at the C-terminal end of somereceptors (Wu et al., 1996) was quickly followed by anx-ray structure of a methyltransferase/Tar–peptidecomplex that demonstrated that such binding oc-curred at a site distant from the methyltransferaseactive site (Djordjevic and Stock, 1998). Whenmounted on this flexible tether, a methyltransferasecan act on the various methylation sites in its own,as well as other nearby, dimers (Li et al., 1997; LeMoual et al., 1997). Continuously exchanging tetra-mers may offer a convenient mechanism for inter-dimer methylation. In solubilized systems such as(Bogonez and Koshland, 1985), the intradimer modi-fication will almost certainly dominate. Both localiz-ing effects, and the ability to methylate multiplesites from the same vantage point, seem to offer clearadvantages. Cellulases, which have similar needs intheir processing of solid cellulose substrates, have astrikingly similar design: a cellulose-binding domainis attached to the catalytic domain via a flexiblelinker (see, e.g., Rouvinen et al., 1990). Not all MCPshave such a methyltransferase-binding tail (Le Moualand Koshland, 1996), but interdimer action allowsthem to be methylated in a fashion reflecting thecellular level of CheA (and thus CheB) activation.The presence of CheA and CheW (i.e., the potentialfor more extensive clustering) did not appear toinfluence the rates of methylation (Le Moual et al.,1997), although it was not proved that clusteringactually occurred under the assay conditions. Theprediction of Bray et al. (1998) that clustering will bevirtually absent at high aspartate concentrationsmay be pertinent.

Recognition of MCP at the methylation sites mayinvolve up to five turns of the helix in which theyreside (Shapiro and Koshland, 1994; Le Moual andKoshland, 1996; Djordjevic and Stock, 1997), al-though the exact mode of binding here is not yetknown. The catalytic site of the methyltransferase iscertainly wide enough to accept a single helix, orperhaps a coiled pair, but may discriminate againstlarger assemblies and thus be affected by the kinase-modulation state of the MCP. That attractant-boundMCPs are inherently better substrates for the meth-yltransferase has been documented in numerousstudies.

Structures of the methylesterase have also been

solved and illustrate how MCP demethylation iscontrolled. A catalytic domain by itself is as active asone in the intact CheR where the associated regula-tory domain has been phosphorylated (Simms et al.,1985; Lupas and Stock, 1989). In the x-ray structureof the intact, unphosphorylated methylesterase(Djordjevic et al., 1998), the active site is physicallyblocked by the regulatory domain, giving rise to thesuggestion that phosphorylation causes the regula-tory domain to physically move out of the way. Anopen methylesterase active site appears broaderthan that of the methyltransferase, which may makeit more accommodating to the ‘‘larger’’ bundle assem-bly of the kinase-activating MCP. That this should bethe only form of the MCP recognized by the methyles-terase has been suggested by the kinetic modelingstudies of Barkai and Leibler (1997).

WORK FOR THE FUTURE

It is obvious from the above discussion that not allquestions have been answered. This is an exciting,but often confusing, time in the field, as manyprevious results need to be reevaluated with newerconcepts in mind. Structures are urgently needed,particularly for the cytoplasmic elements of thesystem and for the partners together in complexes.This is a highly dynamic and complex system, so asingle snapshot of any protein by itself will not tell usall we need to know about its function. It seemsprobable that lower resolution studies such as elec-tron microscopy will produce much of value in thenear future, helping to establish when, and howmany, of the different components are in particularplaces under different conditions. Modeling of thekinetics of the system as a whole is also likely toprove invaluable. The promising starts made in thatdirection (e.g., Bray and Bourret, 1995; Hauri andRoss, 1995; Barkai and Leibler, 1997; Spiro et al.,1997; Levin et al., 1998) have certainly helped intesting whether current assumptions fit the factsand in suggesting new avenues for experimentalwork. Such efforts must go hand in hand, however,with frequent and careful reevaluations of the under-lying assumptions and parameters used.

The authors thank those who assisted us in gathering therecent data, including Ann Stock, Joe Falke, LynnMarie Thomp-son, Gerry Hazelbauer, and Bob Weis. Ida-Maria Sintorn providedinvaluable editorial assistance. Special gratitude is owed to Y.Zhang, P. J. Gardina, J. Christopher, and M. D. Manson of TexasA&M for the model coordinates used to prepare Figure 6.

REFERENCES

Aksamit, R. R., Howlett, B. J., and Koshland, D. E., Jr. (1975)Soluble and membrane-bound aspartate-binding activities inSalmonella typhimurium, J. Bacteriol. 123, 1000–1005.

Alam, M., and Hazelbauer, G. L. (1991) Structural features ofmethyl-accepting taxis proteins conserved between archaebac-

270 MOWBRAY AND SANDGREN

teria and eubacteria revealed by antigenic cross-reaction, J.Bacteriol. 173, 5837–5842.

Alley, M. R. K., Maddock, J. R., and Shapiro, L. (1992) Polarlocalization of a bacterial chemoreceptor, Gene Dev. 6, 825–836.

Alley, M. R. K., Maddock, J. R., and Shapiro, L. (1993) Require-ment of the carboxyl terminus of a bacterial chemoreceptor forits targeted proteolysis, Science 259, 1754–1757.

Ames, P., Chen, J., Wolff, C., and Parkinson, J. S. (1988) Structure–function studies of bacterial chemosensors, Cold Spring HarborSymp. Quant. Biol. 53, 59–65.

Ames, P., and Parkinson, J. S. (1988) Transmembrane signalingby bacterial chemoreceptors: E. coli transducers with lockedsignal output, Cell 55, 817–826.

Ames, P., and Parkinson, J. S. (1994) Constituitively signalingfragments of Tsr, the E. coli serine chemoreceptor, J. Bacteriol.176, 6340–6348.

Ames, P., Yu, Y. A., and Parkinson, J. S. (1996) Methylationsegments are not required for chemotactic signalling by cytoplas-mic fragments of Tsr, the methyl-accepting serine chemorecep-tor of Escherichia coli, Mol. Microbiol. 19, 737–746.

Barkai, N., and Leibler, S. (1997) Robustness in simple biochemi-cal networks [see comments], Nature 387, 913–917.

Barnakov, A. N., Downing, K. H., and Hazelbauer, G. L. (1994)Studies of the structural organization of a bacterial chemorecep-tor by electron microscopy, J. Struct. Biol. 112, 117–124.

Baumgartner, J. W., and Hazelbauer, G. L. (1996) Mutationalanalysis of a transmembrane segment in a bacterial chemorecep-tor, J. Bacteriol. 178, 4651–4660.

Bibikov, S. I., Biran, R., Rudd, K. E., and Parkinson, J. S. (1997) Asignal transducer for aerotaxis in Escherichia coli, J. Bacteriol.179, 4075–4079.

Biemann, H.-P., and Koshland, D. E., Jr. (1994) Aspartate recep-tors of Escherichia coli and Salmonella typhimurium bindligand with negative and half-of-the-sites cooperativity, Biochem-istry 33, 629–634.

Biemann, H. P., Harmer, S. L., and Koshland, D. E., Jr. (1996) Anaspartate/insulin receptor chimera mitogenically activates fibro-blasts, J. Biol. Chem. 271, 27927–27930.

Binnie, R. A., Zhang, H., Mowbray, S. L., and Hermodson, M. A.(1992) Functional mapping of the surface of Escherichia coliribose-binding protein: Mutations that affect chemotaxis andtransport, Protein Sci. 1, 1642–1651.

Bjorkman, A. J., and Mowbray, S. L. (1998) Multiple open forms ofribose-binding protein trace the path of its conformationalchange, J. Mol. Biol. 279, 651–664.

Blair, D. F. (1995) How bacteria sense and swim, Annu. Rev.Microbiol. 49, 489–522.

Bogonez, E., and Koshland, D. E., Jr. (1985) Solubilization of avectorial transmembrane receptor in functional form: Theaspartate receptor of chemotaxis, Proc. Natl. Acad. Sci. USA 82,4891–4895.

Boos, W., and Lucht, J. M. (1996) Periplasmic binding-protein-dependent ABC transporters. in Niedhart, F., Ingraham, R. C.,Lin, E., Low, K., Magasanik, B., Reznikoff, W., Riley, M.,Schaechter, M. & Umbarger, H., (Eds.), Escherichia coli andSalmonella typhimurium: Cellular and Molecular Biology, pp.1175–1209, Am. Soc. Microbiol., Washington, DC.

Boos, W., and Staehelin, A. L. (1981) Ultrastructural localizationof the maltose-binding protein within the cell envelope ofEscherichia coli, Arch. Microbiol. 129, 240–246.

Borkovich, K. A., Alex, L. A., and Simon, M. I. (1992) Attenuationof sensory receptor signaling by covalent modification, Proc.Natl. Acad. Sci. USA 89, 6756–6760.

Borkovich, K. A., Kaplan, N., Hess, J. F., and Simon, M. I. (1989)

Transmembrane signal transduction in bacterial chemotaxisinvolves ligand-dependent activation of phosphate group trans-fer, Proc. Natl. Acad. Sci. USA 86, 1208–1212.

Borkovich, K. A., and Simon, M. I. (1990) The dynamics of proteinphosphorylation in bacterial chemotaxis, Cell 63, 1339–1348.

Bowie, J. U., Pakula, A. A., and Simon, M. I. (1995) The three-dimensional structure of the aspartate receptor from Esch-erichia coli, Acta Crystallogr. D51, 145–154.

Bray, D., and Bourret, R. B. (1995) Computer analysis of thebinding reactions leading to a transmembrane receptor-linkedmultiprotein complex involved in bacterial chemotaxis, Mol.Biol. Cell 6, 1367–1380.

Bray, D., Levin, M. D., and Morton-Firth, C. J. (1998) Receptorclustering as a cellular mechanism to control sensitivity [seecomments], Nature 393, 85–88.

Brooun, A., Zhang, W., and Alam, M. (1997) Primary structure andfunctional analysis of the soluble transducer protein HtrXI inthe archaeon Halobacterium salinarium, J. Bacteriol. 179,2963–2968.

Butler, S. L., and Falke, J. J. (1996) Effects of protein stabilizingagents on thermal backbone motions: A disulfide trappingstudy, Biochemistry 35, 10595–10600.

Butler, S. L., and Falke, J. J. (1998) Cysteine and disulfidescanning reveals two amphiphilic helices in the linker region ofthe asparate chemoreceptor, submitted for publication.

Careaga, C. L., and Falke, J. J. (1992) Structure and dynamics ofEscherichia coli chemosensory receptors: Engineered sulfhy-dryl disulfides, Biophys. J. 62, 209–219.

Chelsky, D., and Dahlquist, F. W. (1980) Chemotaxis in Esch-erichia coli: Associations of protein components, Biochemistry19, 4633–4639.

Chen, X., and Koshland, D. E., Jr. (1995) The N-terminal cytoplas-mic tail of the aspartate receptor is not essential in signaltransduction of bacterial chemotaxis, J. Biol. Chem. 270, 24038–24042.

Chervitz, S. A., and Falke, J. J. (1995) Lock on/off disulfidesidentify the transmembrane signaling helix of the aspartatereceptor, J. Biol. Chem. 270, 24043–24053.

Chervitz, S. A., and Falke, J. J. (1996) Molecular mechanism oftransmembrane signaling by the aspartate receptor: A model,Proc. Natl. Acad. Sci. USA 93, 2545–2550.

Chervitz, S. A., Lin, C. M., and Falke, J. J. (1995) Transmembranesignaling by the aspartate receptor: engineered disulfides re-veal static regions of the subunit interface, Biochemistry 34,9722–9733.

Chothia, C., and Lesk, A. M. (1985) Helix movements in proteins,Trends Biochem. Sci. 10, 116–118.

Clarke, S., and Koshland, D. E., Jr. (1979) Membrane receptors foraspartate and serine in bacterial chemotaxis, J. Biol. Chem.254, 9695–9702.

Cochran, A. G., and Kim, P. S. (1996) Imitation of Escherichia coliaspartate receptor signaling in engineered dimers of the cyto-plasmic domain, Science 271, 1113–1116.

Dahl, M. K., Boos, W., and Manson, M. D. (1989) Evolution ofchemotactic-signal tranducers in enteric bacteria, J. Bacteriol.171, 2361–2371.

Dahl, M. K., and Manson, M. D. (1985) Interspecific reconstitutionof maltose transport and chemotaxis in Escherichia coli withmaltose-binding protein from various enteric bacteria, J. Bacte-riol. 164, 1057–1063.

Danielson, M. A., Bass, R. B., and Falke, J. J. (1997) Cysteine anddisulfide scanning reveals a regulatory alpha-helix in thecytoplasmic domain of the aspartate receptor, J. Biol. Chem.272, 32878–32888.

271CHEMOTAXIS RECEPTORS

Danielson, M. A., Biemann, H.-P., Koshland, D. E., Jr., and Falke,J. J. (1994) Attractant- and disulfide-induced conformationalchanges in the ligand-binding domain of the chemotaxis aspar-tate receptor: A 19F NMR study, Biochemistry 33, 6100–6109.

De Pina, K., Navarro, C., McWalter, L., Boxer, D. H., Price, N. C.,Kelly, S. M., Mandrand-Berthelot, M.-A., and Wu, L.-F. (1995)Purification and characterization of the periplasmic nickel-binding protein NikA of Escherichia coli K12, Eur. J. Biochem.227, 857–865.

Dietzel, I., Kolb, V., and Boos, W. (1978) Pole cap formation inEscherichia coli following induction of the maltose-bindingprotein, Arch. Microbiol. 118, 207–218.

Djordjevic, S., Goudreau, P. N., Xu, Q., Stock, A. M., and West,A. H. (1998) Structural basis for methylesterase CheB regula-tion by a phosphorylation-activated domain, Proc. Natl. Acad.Sci. USA 95, 1381–1386.

Djordjevic, S., and Stock, A. M. (1997) Crystal structure of thechemotaxis receptor methyltransferase CheR suggests a con-served structural motif for binding S-adenosylmethionine, Struc-ture 5, 545–558.

Djordjevic, S., and Stock, A. M. (1998) Chemotaxis receptorrecognition by protein methyltransferase CheR, Nature Struct.Biol. 5, 446–450.

Dunten, P., and Koshland, D. E., Jr. (1991) Tuning the responsive-ness of a sensory receptor via covalent modification, J. Biol.Chem. 266, 1491–1496.

Dunten, P., and Mowbray, S. L. (1995) Crystal structure of thedipeptide binding protein from Escherichia coli involved inactive transport and chemotaxis., Protein Sci. 4, 2327–2334.

Ellis, L., Morgan, D. O., Koshland, D. E., Jr., Clauser, E., Moe,G. R., Bollag, G., Roth, R. A., and Rutter, W. J. (1986) Linkingfunctional domains of the human insulin receptor with thebacterial aspartate receptor, Proc. Natl. Acad. Sci. USA 83,8137–8141.

Falke, J. J., Bass, R. B., Butler, S. L., Chervitz, S. A., andDanielson, M. A. (1997) The two-component signaling pathwayof bacterial chemotaxis: A molecular view of signal transductionby receptors, kinases, and adaptation enzymes, Annu. Rev. CellDev. Biol. 13, 457–512.

Falke, J. J., and Koshland, D. E., Jr. (1987) Global flexibility in asensory receptor: A site-directed cross-linking appproach, Sci-ence 237, 1596–1600.

Flocco, M. M., and Mowbray, S. L. (1994) The 1.9 A x-ray structureof a closed unliganded form of the periplasmic glucose/galactosereceptor from Salmonella typhimurium, J. Biol. Chem. 269,8931–8936.

Flocco, M. M., and Mowbray, S. L. (1998) Conformational changesof periplasmic binding proteins follow a common plan, manu-script in preparation.

Foster, D. L., Mowbray, S. L., Jap, B. K., and Koshland, D. E., Jr.(1985) Purification and characterization of the aspartate chemo-receptor, J. Biol. Chem. 260, 11706–11710.

Gardina, P. J., Bormans, A. F., Hawkins, M. A., Meeker, J. W., andManson, M. D. (1997) Maltose-binding protein interacts simul-taneously and asymmetrically with both subunits of the Tarchemoreceptor, Mol. Microbiol. 23, 1181–1191.

Gardina, P. J., and Manson, M. D. (1996) Attractant signaling byan aspartate chemoreceptor dimer with a single cytoplasmicdomain, Science 274, 425–426.

Gegner, J. A., Graham, D. R., Roth, A. F., and Dahlquist, F. W.(1992) Assembly of an MCP receptor, CheW, and kinase CheAcomplex in the bacterial chemotaxis signal transduction path-way, Cell 70, 975–982.

Grishanin, R. N., and Bibikov, S. I. (1997) Mechanisms of oxygentaxis in bacteria, Biosci. Rep. 17, 77–83.

Hauri, D. C., and Ross, J. (1995) A model of excitation andadaptation in bacterial chemotaxis, Biophys. J. 68, 708–722.

Heldin, C.-H. (1995) Dimerization of cell surface receptors insignal transduction, Cell 80, 213–223.

Higgins, C. F. (1992) ABC transporters: From microorganisms toman, Annu. Rev. Cell Biol. 8, 67–113.

Hughson, A. G., and Hazelbauer, G. L. (1996) Detecting theconformational change of transmembrane signaling in a bacte-rial chemoreceptor by measuring effects on disulfide cross-linking in vivo, Proc. Natl. Acad. Sci. USA 93, 11546–11551.

Hughson, A. G., Lee, G. F., and Hazelbauer, G. L. (1997) Analysisof protein structure in intact cells: Cross-linking in vivo be-tween introduced cyteines in the transmembrane domain of abacterial chemoreceptor, Protein Sci. 6, 315–322.

Iwama, T., Homma, M., and Kawagishi, I. (1997) Uncoupling ofligand-binding affinity of the bacterial serine chemoreceptorfrom methylation- and temperature-modulated signaling states,J. Biol. Chem. 272, 13810–13815.

Jeffery, C. J., and Koshland, D. E., Jr. (1993) Three-dimensionalstructural model of the serine receptor ligand-binding domain,Protein Sci. 2, 559–566.

Kaplan, N., and Simon, M. I. (1988) Purification and characteriza-tion of the wild-type and mutant carboxy-terminal domains ofthe Escherichia coli Tar chemoreceptor, J. Bacteriol. 170,5134–5140.

Kehry, M. R., and Dahlquist, F. W. (1982) The methly-acceptingchemotaxis proteins of E. coli: Identification of the multiplemethylation sites on MCPI, J. Biol. Chem. 257, 10378–10386.

Kolodziej, A. F., Tan, T., and Koshland, D. E., Jr. (1996) Producingpositive, negative, and no cooperativity by mutations at a singleresidue located at the subunit interface in the aspartate recep-tor of Salmonella typhimurium, Biochemistry 35, 14782–14792.

Kossman, M., Wolff, C., and Manson, M. D. (1988) Maltosechemoreceptor of Escherichia coli: Interaction of maltose-binding protein and the tar signal transducer, J. Bacteriol. 170,4516–4521.

Krah, M., Marwan, W., and Oesterhelt, D. (1994) A cytoplasmicdomain is required for the functional interaction of SRI andHtrI in archael signal transduction, FEBS Lett. 353, 301–304.

Lee, G. F., Burrows, G. G., Lebert, M. R., Dutton, D. P., andHazelbauer, G. L. (1994) Deducing the organization of a trans-membrane domain by disulfide cross-linking: The bacterialchemoreceptor Trg, J. Biol. Chem. 269, 29920–29927.

Lee, G. F., Dutton, D. P., and Hazelbauer, G. L. (1995a) Identifica-tion of functionally important helical faces in transmembranesegments by scanning mutagenesis, Proc. Natl. Acad. Sci. USA92, 5416–5420.

Lee, G. F., and Hazelbauer, G. L. (1995) Quantitative approachesto utilizing mutational analysis and disulfide crosslinking formodeling a transmembrane domain, Protein Sci. 4, 1100–1107.

Lee, G. F., Lebert, M. R., Lilly, A. A., and Hazelbauer, G. L. (1995b)Transmembrane signaling characterized in bacterial chemore-ceptors by using sulfhydryl cross-linking in vivo, Proc. Natl.Acad. Sci. USA 92, 3391–3395.

Le Moual, H., and Koshland, D. E., Jr. (1996) Molecular evolutionof the C-terminal cytoplasmic domain of a superfamily ofbacterial receptors involved in taxis, J. Mol. Biol. 261, 568–585.

Le Moual, H., Quang, T., and Koshland, D. E., Jr. (1997) Methyl-ation of the Escherichia coli chemotaxis receptors: Intra- andinterdimer mechanisms, Biochemistry 36, 13441–13448.

Levin, M. D., Morton-Firth, C. J., Abouhamad, W. N., Bourret,R. B., and Bray, D. (1998) Origins of individual swimmingbehavior in bacteria, Biophys. J. 74, 175–181.

Levit, M., Liu, Y., Surette, M. and Stock, J. (1996) Active site

272 MOWBRAY AND SANDGREN

interference and asymmetric activation in the chemotaxis pro-tein histidine kinase CheA, J. Biol. Chem. 271, 32057–32063.

Li, J., Li, G., and Weis, R. M. (1997) The serine chemoreceptorfrom Escherichia coli is methylated through an inter-dimerprocess, Biochemistry 36, 11851–11857.

Lin, L. N., Li, J., Brandts, J. F., and Weis, R. M. (1994) The serinereceptor of bacterial chemotaxis exhibits half-site saturation forserine binding, Biochemistry 33, 6564–6570.

Liu, Y., Levit, M., Lurz, R., Surette, M. G., and Stock, J. B. (1997)Receptor-mediated protein kinase activation and the mecha-nism of transmembrane signaling in bacterial chemotaxis,EMBO J. 16, 7231–7240.

Long, D. G., and Weis, R. M. (1992a) Escherichia coli aspartatereceptor: Oligomerization of the cytoplasmic fragment, Biophys.J. 62, 69–71.

Long, D. G., and Weis, R. M. (1992b) Oligomerization of thecytoplasmic fragment from the aspartate receptor of Esherichiacoli, Biochemistry 31, 9904–9911.

Lupas, A. (1996) Coiled coils: New structures and new functions,Trends Biochem. Sci. 21, 375–382.

Lupas, A., and Stock, J. (1989) Phosphorylation of an N-terminalregulatory domain activates the CheB methylesterase in bacte-rial chemotaxis, J. Biol. Chem. 264, 17337–17342.

Lupas, A., van Dyke, M., and Stock, J. (1991) Predicting coiledcoils from protein sequences, Science 252, 1162–1164.

Lux, R., Jahreis, K., Bettenbrock, K., Parkinson, J. S., andLengeler, J. W. (1995) Coupling the phosphotransferase systemand the methyl-accepting chemotaxis protein-dependent chemo-taxis signaling pathways of Escherichia coli, Proc. Natl. Acad.Sci. USA 92, 11583–11587.

Lynch, B. A., and Koshland, D. E., Jr. (1991) Disulfide cross-linking studies of the transmembrane regions of the aspartatesensory receptor of Escherichia coli, Proc. Natl. Acad. Sci. USA88, 10402–10406.

Maddock, J. R., Alley, M. R., and Shapiro, L. (1993) Polarized cells,polar actions, J. Bacteriol. 175, 7125–7129.

Maddock, J. R., and Shapiro, L. (1993) Polar location of thechemoreceptor complex in the Escherichia coli cell, Science 259,1717–1723.

Manson, M. D., Boos, W., Bassford, P. J., Jr., and Rasmussen, B. A.(1985) Dependence of maltose transport and chemotaxis on theamount of maltose-binding protein, J. Biol. Chem. 260, 9727–9733.

Manson, M. D., and Kossman, M. (1986) Mutations in tar sup-press defects in maltose chemotaxis caused by specific malEmutations, J. Bacteriol. 165, 34–40.

Milburn, M., Prive, G. G., Milligan, D. L., Scott, W. G., Yeh, J.,Jancarik, J., Koshland, D. E., Jr., and Kim, S.-H. (1991)Three-dimensional structures of the ligand-binding domain ofthe bacterial aspartate receptor with and without a ligand,Science 254, 1342–1347.

Milligan, D. L., and Koshland, D. E., Jr. (1988) Site-directedcross-linking: Establishing the dimeric structure of the aspar-tate receptor of bacterial chemotaxis, J. Biol. Chem. 263,6268–6275.

Milligan, D. L., and Koshland, D. E., Jr. (1990) The aminoterminus of the aspartate chemoreceptor is formylmethionine,J. Biol. Chem. 265, 4455–4460.

Milligan, D. L., and Koshland, D. E., Jr. (1991) Intrasubunitsignal transduction by the aspartate chemoreceptor, Science254, 1651–1654.

Milligan, D. L., and Koshland, D. E., Jr. (1993) Purification andcharacterization of the periplasmic domain of the aspartatechemoreceptor, J. Biol. Chem. 268, 19991–19997.

Mizuno, T., and Imae, Y. (1984) Conditional inversion of thethermoresponse in Escherichia coli, J. Bacteriol. 159, 360–367.

Mizuno, T., Mutoh, N., Panasenko, S. M., and Imae, Y. (1986)Acquisition of maltose chemotaxis in Salmonella typhimuriumby the introduction of the Escherichia coli chemosensory trans-ducer gene, J. Bacteriol. 165, 890–895.

Moe, G. R., Bollag, G. E., and Koshland, D. E., Jr. (1989)Transmembrane signaling by a chimera of the Escherichia coliaspartate receptor and the human insulin receptor, Proc. Natl.Acad. Sci. USA 86, 5683–5687.

Moe, G. R., and Koshland, D. E., Jr. (1986). Transmembranesignaling through the aspartate receptor. in Youvan, D. C., andDaldal, F. (Eds.), Microbiology Energy Transduction: Genetics,Structure and Function of Membrane Proteins, pp. 163–176,Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY.

Morgan, D. G., Baumgartner, J. W., and Hazelbauer, G. L. (1993)Proteins antigenically related to methyl-accepting chemotaxisprotein of Escherichia coli detected in a wide range of bacterialspecies, J. Bacteriol. 175, 133–140.

Mowbray, S. L., and Cole, L. B. (1992) 1.7 A X-ray structure of theperiplasmic ribose receptor from Escherichia coli, J. Mol. Biol.225, 155–175.

Mowbray, S. L., Foster, D. L., and Koshland, D. E., Jr. (1985)Proteolytic fragments identified with domains of the aspartatechemoreceptor, J. Biol. Chem. 260, 11711–11718.

Mowbray, S. L., and Koshland, D. E., Jr. (1987) Additive andindependent responses in a single receptor: Aspartate andmaltose stimuli on the Tar protein, Cell 50, 171–180.

Mutoh, N., Oosawa, K., and Simon, M. I. (1986) Characterizationof Escherichia coli chemotaxis receptor mutants with nullphenotypes, J. Bacteriol. 167, 992–998.

Nathan, P., Gomes, S. L., Hahnenberger, K., Newton, A., andShapiro, L. (1986) Differential localization of membrane recep-tor chemotaxis proteins in the Caulobacter predivisional cell, J.Mol. Biol. 191, 433–440.

Nickitenko, A. V., Trakhanov, S., and Quiocho, F. A. (1995) 2 Aresolution structure of DppA, a periplasmic dipeptide transport/chemosensory receptor, Biochemistry 34, 16585–16595.

Ninfa, A. J., Ninfa, E. G., Lupas, A. N., Stock, A., Magasanik, B.,and Stock, J. (1988) Crosstalk between bacterial chemotaxissignal transduction proteins and regulators of transcription ofthe Ntr regulon: Evidence that nitrogen assimilation andchemotaxis are controlled by a common phosphotransfer mech-anism, Proc. Natl. Acad. Sci. USA 85, 5492–5496.

Ninfa, E. G., Stock, A., Mowbray, S. L., and Stock, J. (1991)Reconstitution of the bacterial chemotaxis signal transductionsystem from purified components, J. Biol. Chem. 266, 9764–9770.

Nowlin, D. M., Bollinger, J., and Hazelbauer, G. L. (1987) Sites ofcovalent modification in Trg, a sensory transducer of Esch-erichia coli, J. Biol. Chem. 262, 6039–6045.

Oosawa, K., Mutoh, N., and Simon, M. I. (1988) Cloning of theC-terminal cytoplasmic fragment of the tar protein and effectsof the fragment on chemotaxis of Escherichia coli, J. Bacteriol.170, 2521–2526.

Ottemann, K. M., and Koshland, D. E., Jr. (1997) Converting atransmembrane receptor to a soluble receptor: recognitiondomain to effector domain signaling after excision of the trans-membrane domain, Proc. Natl. Acad. Sci. USA 94, 11201–11204.

Ottemann, K. M., Thorgeirsson, T. E., Kolodziej, A. F., Shin, Y. K.,and Koshland, D. E., Jr. (1998) Direct measurement of smallligand-induced conformational changes in the aspartate chemo-receptor using EPR, Biochemistry 37, 7062–7069.

Pakula,A.A., and Simon, M. I. (1992) Determination of transmem-

273CHEMOTAXIS RECEPTORS

brane protein structure by disulfide cross-linking: The Esch-erichia coli Tar receptor, Proc. Natl. Acad. Sci. USA 89,4144–4148.

Quiocho, F. A., and Ledvina, P. S. (1996) Atomic structure andspecificity of bacterial periplasmic receptors for active transportand chemotaxis: Variation of common themes, Mol. Microbiol.20, 17–25.

Reader, R. W., Tso, W. W., Springer, M. S., Goy, M. F., and Adler, J.(1979) Pleiotropic aspartate taxis and serine taxis mutants ofEscherichia coli, J. Gen. Microbiol. 111, 363–374.

Rebbapragada, A., Johnson, M. S., Harding, G. P., Zuccarelli, A. J.,Fletcher, H. M., Zhulin, I. B., and Taylor, B. L. (1997) The Aerprotein and the serine chemoreceptor Tsr independently senseintracellular energy levels and transduce oxygen, redox, andenergy signals for Escherichia coli behavior, Proc. Natl. Acad.Sci. USA 94, 10541–10546.

Rice, M. S., and Dahlquist, F. W. (1991) Sites of deamidation andmethylation in Tsr, a bacterial chemotaxis sensory transducer,J. Biol. Chem. 266, 9746–9753.

Rollins, C., and Dahlquist, F. W. (1981) The methyl-acceptingchemotaxis proteins of E. coli: A repellent-stimulated, covalentmodification, distinct from methylation, Cell 25, 333–340.

Rouvinen, J., Bergfors, T., Teeri, T., Knowles, J. K. C., and Jones,T. A. (1990) Three-dimensional structure of cellobiohydrolase IIfrom Trichoderma reesei, Science 249, 380–386.

Sack, J. S., Saper, M. A., and Quiocho, F. A. (1989) Periplasmicbinding protein structure and function: refined X-ray structuresof the leucine/isoleucine/valine-binding protein and its complexwith leucine, J. Mol. Biol. 206, 171–191.

Schuster, S. C., Swanson, R. V., Alex, L. A., Bourret, R. B., andSimon, M. I. (1993) Assembly and function of a quaternarysignal transduction complex monitored by surface plasmonresonance, Nature 365, 343–347.

Scott, W. G., and Stoddard, B. L. (1994) Transmembrane signal-ling and the aspartate receptor, Structure 2, 877–887.

Seeley, S. K., Weis, R. M., and Thompson, L. K. (1996) Thecytoplasmic fragment of the aspartate receptor displays glo-bally dynamic behavior, Biochemistry 35, 5199–5206.

Segall, J. E., Block, S. M., and Berg, H. C. (1986) Temporalcomparisons in bacterial chemotaxis, Proc. Natl. Acad. Sci.USA 83, 8987–8991.

Shapiro, M. J., and Koshland, D. E., Jr. (1994) Mutagenic studiesof the interaction between the aspartate receptor and methyl-transferase from Escherichia coli, J. Biol. Chem. 269, 11054–11059.

Sharff, A. J., Rodseth, L. E., Spurlino, J. C., and Quiocho, F. A.(1992) Crystallographic evidence of a large ligand-inducedhinge-twist motion between the two domains of the maltodex-trin binding protein involved in active transport and chemo-taxis, Biochemistry 31, 10657–10663.

Sherris, D., and Parkinson, J. S. (1981) Posttranslational process-ing of methyl-accepting chemotaxis proteins in Escherichia coli,Proc. Natl. Acad. Sci. USA 78, 6051–6055.

Shilton, B. H., Flocco, M. M., Nilsson, M., and Mowbray, S. L.(1996) Conformational changes of three periplasmic receptorsfor bacterial chemotaxis and transport: The maltose-, glucose/galactose- and ribose-binding proteins, J. Mol. Biol. 264, 350–363.

Simms, S. A., Keane, M. G., and Stock, J. (1985) Multiple forms ofthe CheB methylesterase in bacterial chemosensing, J. Biol.Chem. 260, 10161–10168.

Spiro, P. A., Parkinson, J. S., and Othmer, H. G. (1997) A model ofexcitation and adaptation in bacterial chemotaxis, Proc. Natl.Acad. Sci. USA 94, 7263–7268.

Springer, M. S., Goy, M. F., and Adler, J. (1979) Protein methyl-ation in behavioural control mechanisms and in signal transduc-tion, Nature 280, 279–284.

Springer, W. R., and Koshland, D. E., Jr. (1977) Identification of aprotein methyltransferase as the cheR gene product in thebacterial sensing system, Proc. Natl. Acad. Sci. USA 74, 533–537.

Spudich, J. L. (1994) Protein–protein interaction converts aproton pump into a sensory receptor, Cell 79, 747–750.

Spurlino, J. C., Lu, G.-Y., and Quiocho, F. A. (1991) The 2.3 Aresolution structure of the maltose- or maltodextrin-bindingprotein, a primary receptor of bacterial active transport andchemotaxis, J. Biol. Chem. 266, 5202–5219.

Stock, A. M., and Mowbray, S. L. (1995) Bacterial chemotaxis: afield in motion, Curr. Opin. Struct. Biol. 5, 744–751.

Stock, J. B., Lukat, G., and Stock, A. M. (1991) Bacterial chemo-taxis and the molecular logic of intracellular signal transduc-tion networks, Annu. Rev. Biophys. Biophys. Chem. 20, 109–136.

Stock, J. B., and Surette, M. G. (1996). Chemotaxis. in Niedhart,F., Ingraham, R. C., Lin, E., Low, K., Magasanik, B., Reznikoff,W., Riley, M., Schaechter, M., and Umbarger, H. (Eds.), Esch-erichia coli and Salmonella typhimurium: Cellular and Molecu-lar Biology, 2nd ed., pp. 1103–1129, Am. Soc. Microbiol. Press,Washington, DC.

Stock, J. B., Surette, M. G., Levit, M., and Park, P. (1995).Two-component signal transduction systems: structure func-tion relationships and mechanisms of catalysis. in Hoch, J. A.,and Silhavy, T. J. (Eds.), Two-Component Signal Transduction,(pp. 25–51, Am. Soc. Microbiol., Washington, DC.

Stoddard, B. L., Bui, J. D., and Koshland, D. E., Jr. (1992)Structure and dynamics of transmembrane signaling by theEscherichia coli aspartate receptor, Biochemistry 31, 11978–11983.

Stoddard, B. L., and Koshland, D. E., Jr. (1992) Prediction of thestructure of a receptor–protein complex using a binary dockingmethod, Nature 358, 774–776.

Stoddard, B. L., and Koshland, D. E., Jr. (1993) Molecularrecognition analyzed by docking simulations: The aspartatereceptor and isocitrate dehydrogenase from Escherichia coli,Proc. Natl. Acad. Sci. USA 90, 1146–1153.

Surette, M., and Stock, J. B. (1996) Role of a-helical coiled-coilinteractions in receptor dimerization, signaling, and adaptationduring bacterial chemotaxis, J. Biol. Chem. 271, 17966–17973.

Surette, M. G., Levit, M., Liu, Y., Lukat, G., Ninfa, E., Ninfa, A.,and Stock, J. B. (1996) Dimerization is required for the activityof the protein histidine kinase CheA that mediates signaltransduction in bacterial chemotaxis, J. Biol. Chem. 271, 939–945.

Tam, R., and Saier, M. H., Jr. (1993) Structural, functional, andevolutionary relationships among extracellular solute-bindingreceptors of bacteria, Microbiol. Rev. 57, 320–346.

Tatsuno, I., Homma, M., Oosawa, K., and Kawagishi, I. (1996)Signaling by the Escherichia coli aspartate chemoreceptor Tarwith a single cytoplasmic domain per dimer [see comments],Science 274, 423–425.

Terwilliger, T. C., Bogonez, E., Wang, E. A., and Koshland, D. E.,Jr. (1983) Sites of methyl esterification on the aspartate recep-tor involved in bacterial chemotaxis, J. Biol. Chem. 258,9608–9611.

Terwilliger, T. C., and Koshland, D. E., Jr. (1984) Sites of methylesterification and deamidation on the aspartate receptor in-volved in chemotaxis, J. Biol. Chem. 259, 7719–7725.

Wang, J., Balazs, Y. S., and Thompson, L. K. (1997) Solid-state

274 MOWBRAY AND SANDGREN

REDOR NMR distance measurements at the ligand site of abacterial chemotaxis membrane receptor, Biochemistry 36, 1699–1703.

Wolff, C. (1983) Genetic and biochemical studies of maltosechemotaxis in Escherichia coli, Diplom Thesis, Department ofBiology, University of Konstanz, Konstanz, West Germany.

Wu, J., Li, J., Li, G., Long, D. G., and Weis, R. M. (1996) Thereceptor binding site for the methyltransferase of bacterialchemotaxis is distinct from the sites of methylation, Biochemis-try 35, 4984–4993.

Wu, J., Long, D. G., and Weis, R. M. (1995) Reversible dissociationand unfolding of the Escherichia coli aspartate receptor cytoplas-mic fragment, Biochemistry 34, 3056–3065.

Wurgler-Murphy, S. M., and Saito, H. (1997) Two-componentsignal transducers and MAPK cascades, Trends Biochem. Sci.22, 172–176.

Yaghmai, R., and Hazelbauer, G. L. (1992) Ligand occupancymimicked by single residue substitutions in a receptor: Trans-membrane signalling induced by mutation, Proc. Natl. Acad.Sci. USA 89, 7890–7894.

Yamamoto, K., and Imae, Y. (1993) Cloning and characterizationof the Salmonella typhimurium-specific chemoreceptor Tcp fortaxis to citrate and from phenol, Proc. Natl. Acad. Sci. USA 90,217–221.

Yeh, J. I., Biemann, H.-P., Pandit, J., Koshland, D. E., Jr., andKim, S.-H. (1993) The three-dimensional structure of the ligand-binding domain of a wild-type bacterial chemotaxis receptor, J.Biol. Chem. 268, 9787–9792.

Yeh, J. I., Biemann, H. P., Prive, G. G., Pandit, J., Koshland, D. E.,Jr., and Kim, S. H. (1996) High-resolution structures of theligand binding domain of the wild-type bacterial aspartatereceptor, J. Mol. Biol. 262, 186–201.

Zhang, W. A., Brooun, J., MacCandless, J., Banda, P., and Alam,M. (1996) Signal transduction in the archaeon Halobacteriumsalinarium is processed through three subfamilies of 13 solubleand membrane-bound transducer proteins, Proc. Natl. Acad.Sci. USA 93, 4649–4654.

Zhang, Y., Conway, C., Rosato, M., Suh, Y., and Manson, M. (1992)Maltose chemotaxis involves residues on the same face of theN-terminal and C-terminal domains of maltose-binding protein,J. Biol. Chem. 267, 22813–22820.

Zhang, Y., Gardina, P. J., Christopher, J., and Manson, M. D.(1998) Model of maltose-binding protein/chemoreceptor com-plex supports intrasubunit signaling mechanism, Proc. Natl.Acad. Sci. USA, submitted for publication.

Zhulin, I. B., Johnson, M. S., and Taylor, B. L. (1997) How dobacteria avoid high oxygen concentrations? Biosci. Rep. 17,335–342. imefor a

275CHEMOTAXIS RECEPTORS