PhoB Homolog from Thermotoga maritima

PhoB Homolog from Thermotoga maritima

Structure, Vol. 10, 153–164, February, 2002, 2002 Elsevier Science Ltd. All rights reserved. PII S0969-2126(01)00706-7 Evidence of Intradomain and ...

705KB Sizes 0 Downloads 44 Views

Structure, Vol. 10, 153–164, February, 2002, 2002 Elsevier Science Ltd. All rights reserved.

PII S0969-2126(01)00706-7

Evidence of Intradomain and Interdomain Flexibility in an OmpR/PhoB Homolog from Thermotoga maritima David R. Buckler,1 Yuchen Zhou,2,4 and Ann M. Stock1,2,3 Center for Advanced Biotechnology and Medicine 679 Hoes Lane Piscataway, New Jersey 08854 1 Department of Biochemistry University of Medicine and Dentistry of New Jersey Robert Wood Johnson Medical School Piscataway, New Jersey 08854 2 Howard Hughes Medical Institute

Summary Two-component systems, the predominant signal transduction strategy used by prokaryotes, involve phosphorelay from a sensor histidine kinase (HK) to an intracellular response regulator protein (RR) that typically acts as a transcription regulator. RRs are modular proteins, usually composed of a conserved regulatory domain, which functions as a phosphorylation-activated switch, and an attached DNA binding effector domain. The crystal structure of a Thermotoga maritima transcription factor, DrrD, has been determined at 1.5 A˚ resolution, providing the first structural information for a full-length member of the OmpR/ PhoB subfamily of RRs. A small interdomain interface occurs between ␣5 of the regulatory domain and an antiparallel sheet of the effector domain. The lack of an extensive interface in the unphosphorylated protein distinguishes DrrD from other structurally characterized multidomain RRs and suggests a different mode of interdomain regulation. Introduction Sensing and responding to the environment is a critical survival strategy used by prokaryotes. Such behavior is often mediated by so-called two-component signaling systems [1]. These systems are prevalent in microorganisms but appear to be rare or absent in animals and, hence, present attractive targets for the development of antimicrobial agents. In typical two-component signaling pathways, the first component is a sensor histidine kinase (HK), usually anchored to the inner cell membrane, which undergoes ATP-dependent autophosphorylation at a conserved histidine residue. The resulting high-energy phosphoryl group is then transferred to a conserved aspartyl residue of a second component of the pathway, the intracellular response regulator protein (RR). RRs most commonly function as transcription factors, and phosphorylation typically enhances their binding to target sites on DNA. The intracellular level of phosphorylated RR is regulated by the HK, which not only donates a phosphoryl group to the RR 3 4

Correspondence: [email protected] Present Address: Merck & Co., Inc., Rahway, New Jersey 07065.

but may also serve as a phosphatase of the phosphorylated RR. The ratio of kinase to phosphatase activity for a given HK is in many cases modulated by the interaction of specific environmental stimuli with a sensor domain in the HK, thus providing the link between environmental cue and gene expression, mediated by intracellular levels of phosphorylated RR. Both HK and RR signaling proteins exhibit modular architectures, highlighted by conserved residues critical in the biochemistry of phosphoryl transfer [2]. The defining regulatory domain, shared by all RRs, serves as a phosphorylation-activated switch and has been the subject of extensive structural studies in both activated and nonactivated forms. Typically, RRs are two-domain proteins in which the N-terminal regulatory domain is followed by a C-terminal effector domain harboring a helix-turn-helix DNA binding motif. RRs subdivide into three major subfamilies of transcription regulators based on effector domain sequence. Crystal or solution structures of the effector domains from the three major subfamilies, the OmpR/PhoB [3–5], the NarL/FixJ [6, 7], and the NtrC [8] subfamilies, have been determined; however, the full-length structure of only the NarL/FixJ subfamily of transcription regulators has been solved to date [6, 7]. We report here the first structure of a fulllength OmpR/PhoB subfamily member determined at 1.50 A˚ resolution from a crystal of DrrD, a homolog from Thermotoga maritima. The OmpR/PhoB subfamily is the largest single subfamily of RRs, based on sequence database searches and full genome analyses [9–12], and includes many well-studied and biomedically relevant members. OmpR was initially identified in Escherichia coli as a transcription regulator governing outer membrane porin expression in response to osmotic stress [1]. Recent studies on biofilm formation in E. coli indicate that OmpR [13] together with a second OmpR/PhoB subfamily member, CpxR [14], are involved in biofilm formation in pathogenic strains of this species. Other notable members of this subfamily include: VanR, involved in regulation of vancomycin resistance genes in Enterococcus (hospital-acquired infection) [15]; VncR, responsible for repressing antibiotic-induced cell lysis in the important pathogen Streptococcus pneumoniae (bacterial meningitis, pneumonia, and otitis media) [16]; PhoB and PhoP, involved in phosphate assimilation and metabolism in E. coli [17] and Bacillus subtilis [18], respectively; and PmrA, involved in polymyxin resistance in Salmonella [19]. Recently, deletion of two-component systems involving OmpR/PhoB subfamily members in S. pneumoniae has identified pathways that are either essential to cell viability or important in virulence in in vivo models for respiratory infection [11, 12]. Despite the availability of extensive information on the isolated domains, the specific structural details of how phosphorylation of a regulatory domain modulates the Key words: OmpR; PhoB; response regulator; two component; winged helix

Structure 154

function of an attached effector domain remains largely unanswered and will require structural data from additional multidomain RRs in phosphorylated and nonphosphorylated forms and those in complex with target DNA and elements of the transcription machinery. For the two previously solved structures of two-domain RRs, transcription regulator NarL from E. coli [6] and the receptor modification enzyme methylesterase CheB from Salmonella typhimurium [20], an extensive interdomain interface positions the regulatory domain near key functional regions of the effector domain, providing steric inhibition in the nonphosphorylated state. DrrD, in contrast, exhibits a much smaller interdomain interface, and the interdomain linker is largely disordered in the crystal lattice. This suggests a system lacking strict enforcement of a specific interdomain orientation in the nonphosphorylated form. Additionally, a novel orientation of ␣4 in the regulatory domain has been observed, suggestive of conformational mobility for a key helical segment that has been previously implicated in phosphorylation-induced conformational transitions. These overall features and the functional relevance of implied conformational flexibility within and between domains are presented here.

Results Identification of OmpR/PhoB Homologs in T. maritima Extensive attempts to crystallize full-length OmpR from E. coli have failed. Thus, we pursued OmpR/PhoB homologs from the hyperthermophile T. maritima, for which genome sequence analysis revealed 11 open reading frames containing RR regulatory domains [21]. A first OmpR/PhoB homolog from T. maritima, DrrA (DNA binding response regulator A), was discovered prior to the availability of the genome sequence [22]; we identified three additional OmpR/PhoB subfamily members through sequence analysis of the effector regions. The new homologs identified here are denoted DrrB, DrrC, and DrrD and correspond to TIGR annotations [21] BTM126, BTM1259, and BTM399, respectively; DrrA is BTM1655. The range of pairwise sequence identities within the group of four T. maritima OmpR/PhoB homologs (32%–40% for the six comparisons) was similar to that between E. coli OmpR and the four thermophilic proteins (30%–37% for the four comparisons). Each of the three new OmpR/PhoB homologs was overexpressed in E. coli, purified, and subjected to crystallization screens; single crystals affording X-ray diffraction to 1.50 A˚ were obtained for the DrrD homolog, whose structure is reported in this work.

Structure Determination and Overall Model Quality Attempts to produce initial phase estimates by molecular replacement using alanine models of either multiple regulatory domain structures or of the OmpR DNA binding domain failed. Therefore, we carried out a three wavelength anomalous dispersion experiment using crystals of a selenomethionine derivative of DrrD, which

Table 1. Data Collection Statistics Resolution range (A˚) (outer shell) Rmerge, all data (outer shell) Rmeasa, all data (outer shell) Total reflections (outer shell) Multiplicity, all data (outer shell) Completeness, all data (%) (outer shell) I/sigI, all data (outer shell) a

20.0–1.50 1.58–1.50 0.052 0.139 0.060 0.170 40,466 5891 5.4 5.4 99.6 100.0 7.5 5.0

[69]

produced X-ray diffraction data to 1.50 A˚ (Table 1). The space group is P21 (cell constants: a ⫽ 34.57 A˚, b ⫽ 71.30 A˚, c ⫽ 54.61 A˚, and ␤ ⫽ 106.6⬚), and the asymmetric unit contains one molecule with a Matthews coefficient of 2.5 (solvent content 50%). The current model, shown in Figure 1, was refined to Rfree ⫽ 0.210 (Rcryst ⫽ 0.179) using diffraction data extending to 1.50 A˚ Bragg spacing (Table 2). The model contains 206 water oxygens and six thiocyanate (SCN) ions. The SCN anions were typically associated with surface-exposed Arg or Lys residues, with the exception of one SCN, which was buried in the interdomain interface. Lack of well-defined electron density in two regions (the interdomain linker from 124– 125 and the interhelical helix-turn-helix loop in the DNA binding domain from 182–187) prevented complete model building through these regions. A limited interdomain interface that involved well-ordered residues was observed (Figure 1); some lattice interactions also occur in the vicinity of the interface as discussed separately below. Of the modeled residues, analysis of main chain dihedral angles by Procheck [23] showed 93.8% in most favored regions, 5.2% in additionally allowed regions, one in a generously allowed region, and one (V58) in a disallowed region. Interestingly, V58 occurs in the loop extending from the site of phosphorylation at the tip of ␤3; in other regulatory domains, this position is often Gly, which would permit this main chain conformation at much lower energetic cost. Average atomic B factors for main chain atoms, side chain atoms, all protein atoms, and water atoms are 15, 17, 16, and 30 A˚2, respectively. The overall G factor [23] for the current model is 0.20. The Regulatory Domain The overall fold of the regulatory domain is similar to that of other regulatory domains of the RR superfamily [24]. The topology is an ␣/␤ sandwich, with a central five-stranded parallel sheet surrounded by two ␣ helices on one face and three on the other. From both biochemical and structural studies of members of the RR superfamily, it has been established that phosphorylation occurs at an absolutely conserved aspartate positioned at the tip of the centermost ␤ strand (␤3), and phosphoryl transfer from phosphodonor to aspartyl carboxylate is promoted through autocatalysis, dependent on Mg(II),

Figure 1. Structure of Full-Length DrrD (A) A ribbon representation of the DrrD structure highlighting residues at the interdomain interface (ball and stick rendering) and six thiocyanate ions (CPK rendering). For the stick and CPK renderings, carbon is in green, nitrogen is in blue, oxygen is in red, and sulfur is in yellow. Secondary structural elements are labeled ␣1–5 and ␤1–5 for the regulatory domain and H1–3 and S1–7 for the effector domain. The recognition helix, H3, in the effector domain is in gold. Two regions through which electron density was not observed, the interdomain linker and the loop connecting H2 and H3, are indicated by dashed lines. (B) Stereo C␣ trace rendered in the same orientation as in (A). This figure and the protein models shown in Figures 3–7 were generated using Ribbons [65] Version 3.14.

Structure 156

Table 2. Refinement Statistics Resolution range (A˚) Rcryst Rfree Number of residues Missing residues Number of atoms Protein Solvent SCN Rmsd bonds (A˚) Rms angles (⬚) Average B factor (A˚2)

20.0–1.50 0.179 0.210 224 8

Main chain Side chain Solvent SCN Ramachandran plot Most favored (%) Additional allowed (%) Generously allowed (%) Disallowed (%)

14.9 17.3 29.6 35.7

1792 206 18 0.020 1.65

93.8 5.2 0.5 0.5

which is bound near the site of phosphorylation by additional conserved acidic residues. A structure-based sequence alignment [25] of DrrD and several representative regulatory domains is shown in Figure 2. In this report, regulatory domain helices and strands are referred to as ␣N and ␤N (N ⫽ 1–5), respectively; those of the effector domain are referred to as HN (N ⫽ 1–3) and SN (N ⫽ 1–7), respectively, to conform with numbering defined in earlier studies of the isolated domains. Site of Phosphorylation and Functionally Conserved Residues As in other regulatory domain structures, an acidic pocket formed by conserved residues surrounds the phosphoryl acceptor, D53 (Figure 3), and a water oxygen with low B factor (W1, 13 A˚2) occupies the acidic pocket. Lys-103 corresponds to a highly conserved residue at the C terminus of ␤5 and extends toward the acidic pocket (Figure 3), as has been found in previously described regulatory domain structures. It is followed by a highly conserved proline, and this KP peptide bond in DrrD is in the cis conformation, as has been observed for previously reported regulatory domain structures excluding methylesterase CheB [20]. The conserved aromatic residue on ␤5 corresponding to Y100 in DrrD has been observed in various rotameric states in the different regulatory domain structures and in both “outward” and “inward” conformations in the E. coli CheY structure solved at 1.7 A˚ [26]. The DrrD Y100 side chain appears to be stabilized by significant ␲-␲ interactions between the Y100 aromatic ring and the D85 carboxylate group; this also shields the Y100 side chain from exposure to bulk solvent. Additional stabilizing interactions for this Y100 conformation include bifurcated H-bonds to two main chain NHs in the ␣4-␤5 loop. This stabilized, “inward” orientation of Y100 positions the side chain adjacent to that of the conserved residue T81 on ␤4 (Figure 3). Repositioning of T81, expected upon phosphorylation, would therefore promote Y100 to adopt alternate orientations as part of the general activation mechanism, suggested from analysis of activated regulatory domains for which these key conserved

side chains undergo linked movements following phosphorylation [27–32] (for reviews, see [33–35]). Regulatory Domain ␣4 The most striking feature of the DrrD regulatory domain relative to others is the ␣4 helix, a region that undergoes significant structural changes upon activation in structures of isolated regulatory domains [27–32]. For DrrD, this helical segment adopts a more perpendicular arrangement relative to the central ␤ strand in comparison with other regulatory domain structures, as shown in Figure 4 for DrrD and PhoBN [36]. The different orientations arise from the variations in backbone dihedral angles highlighted in the traced Ramachandran plots shown in Figure 4. With three exceptions, the residues in the 82–94 region of DrrD and the corresponding 84–96 region of PhoBN occupy similar positions in the Ramachandran plot. These differences in backbone torsion angles result in a relative rotation of the ␣4 helical segment both about its own axis and perpendicular to the midpoint of its axis when comparing DrrD and PhoB (Figure 4). Specific hydrophobic interactions appear to stabilize these alternative orientations. As described previously [36], the orientation of ␣4 in the PhoBN crystal structure is anchored to the protein core primarily by the side chain of a single hydrophobic residue, V92, which occupies a hydrophobic pocket created by W54 (␤3-␣3 loop), T83 (␤4), T97 (␣4-␤5 loop), and Y102 (␤5), as shown in Figure 4. This PhoBN-␣4 anchor residue is highly conserved as hydrophobic in the multiple sequence alignment of OmpR/PhoB regulatory domains (Figure 2). Interestingly, despite the occurrence of Val in this position in DrrD (V90), the side chain of an alternate hydrophobic residue within ␣4, Y88, provides the primary hydrophobic anchor to its protein core in a fashion analogous to that of PhoB V92 (Figure 4). The pocket is formed in DrrD by I54 (␤3-␣3 loop), T81 (␤4), M95 (␣4-␤5 loop), and Y100 (␤5) residues corresponding in both the superposed structures and in the multiple sequence alignment position to those described above for the PhoBN hydrophobic pocket (Figure 4; see also Figure 2). Thus, it appears that an analogous hydrophobic region within the regulatory domain of the two OmpR/PhoB subfamily members provides the anchoring region for the helix in both cases, but a shift in register by two residues of the ␣4 anchor residue together with the changes in dihedral angles noted above through the ␤4-␣4 region result in different orientations for ␣4. The Effector Domain The structures of the isolated effector domains of OmpR in the crystal [3, 4] and PhoB in solution [5] have been described, and a structure-based sequence alignment [25] of three of these is shown in Figure 5. The fold belongs to the winged-helix subfamily [37] but is differentiated by having an N-terminal four-stranded antiparallel sheet and an unusually long loop connecting H2 and H3 of the helix-turn-helix motif of this domain [38]. The largest backbone deviations for these three structures occur in the loops extending from both ends of the recognition helix, H3, as was also noted in the comparison of the PhoB and OmpR DNA binding domains [5]. The H2-H3 loop of DrrD is N-terminal to the recogni-

Full-Length OmpR Homolog from T. maritima 157

Figure 2. Sequence Alignment of DrrD and Representative, Structurally Characterized Response Regulators Structure-based sequence alignment for DrrD and isolated response regulator (RR) regulatory domains and effector domains was performed using STAMPS [25] and rendered with ALSCRIPT [66]. The structures compared, including their Protein Data Bank codes, the overall C␣ rms difference from DrrD, and the number of equivalent residues used in the rms difference calculation in parenthesis, are as follows: PhoBN (1B00 [36], 1.36 A˚, 99), FixJN (1DCK [67], 1.48 A˚, 96), CheY (3CHY [26], 1.78 A˚, 102), and NarL (1A04 [6], 2.16 A˚, 76) for regulatory domains and OmpRC (1OPC [3],1.7 A˚, 73) and PhoBC (1QQI [5], 1.3 A˚, 72) for effector domains. Strands and helices, as defined using DSSP [68], are shaded with black and gray backgrounds, respectively. The consensus residues (⬎50% occurrence) for each position in a 269 member multiple sequence alignment of OmpR/PhoB subfamily members appear below the sequence alignment; dashes indicate no residue with occurrence frequency ⬎50%. A schematic diagram of secondary structural elements for DrrD appears below the consensus sequence using the nomenclature described in Figure 1. A histogram below the sequences indicates lattice contacts (assessed using CCP4 Areaimol calculation of solventaccessible surface area protected by lattice contacts, y range ⫽ 0–168 A˚2). Residues that are involved in the interdomain interface are indicated with open triangles; residues in the DrrD sequence for which lack of electron density prevented model building are shown in reduced font size. The numbering at the top corresponds to the DrrD sequence.

tion helix and includes residues L180–V187. This loop exhibits clear conformational disorder, as electron density for only two of these residues, L180 and W181, is discernible in the electron density maps. The H3-S6 loop (Figures 2 and 5) extends from the C-terminal end of the recognition helix and includes K205-I211 in DrrD, E213-Y221 in OmpRC [3], and A82-M89 in PhoBC [5]. The

differing lengths of these loops arise from both a singleresidue deletion for OmpR and different relative lengths of the H3 helix (Figure 2). Interdomain Interactions This model of DrrD provides the first high-resolution structural data of interdomain contacts occurring be-

Structure 158

Figure 3. Regulatory Domain of DrrD Highlighting Active Site and Key Conserved Residues Functionally conserved residues are rendered in ball and stick (carbon in black, oxygen in red, and nitrogen in blue) and surrounded by electron density maps generated with 2Fo ⫺ 2Fc Fourier coefficients (contoured at 1.0 ␴ and cut with a 2.0 A˚ tolerance around side chain atoms). The site of phosphorylation, D53, and the key conserved residues, T81, Y100, and K103, that facilitate phosphorylationinduced conformational change are labeled.

tween regulatory and effector domains for a member of the OmpR/PhoB subfamily. The interaction is mediated by the C-terminal half of ␣5 of the regulatory domain and the conserved four-stranded antiparallel sheet at the N terminus of the effector domain. The individual residues contributing to this interface, assessed by degree of surface area protected from solvent accessibility in the full-length model relative to isolated domains, are

(Figures 1, 2, and 6): R113 (4 A˚2), A116 (26 A˚2), R119 (110 A˚2), R120 (68 A˚2), and S122 (28 A˚2) in the regulatory domain and K128 (65 A˚2), D133 (10 A˚2), I135 (49 A˚2), D137 (3 A˚2), Y144 (47 A˚2), G146 (9 A˚2), and S147 (70 A˚2) in the effector domain. Of note, electron density maps of the interdomain region contained distinct contours of bilobal but continuous electron density (resembling an intact peanut shell), which could be attributed to a thiocyanate ion (SCN). As observed in other high-resolution structures, the sp-hybridized CN moiety appears nearly spherical and comparable in radius to the S moiety of the thiocyanate ion, resulting in a bilobal appearance in electron density maps. This interfacial SCN anion was completely buried from solvent, occupying a partially hydrophobic, partially positively charged pocket delimited by side chains of A116, R119, and R120 from the regulatory domain and Y144 and I135 from the effector domain. The total interaction surface between regulatory and effector domains for DrrD is small, amounting to only 245 A˚2. In addition, much of the segment constituting the interdomain linker (S121-K127) did not exhibit well-defined electron density, indicating disorder or multiple discrete conformations in the crystal lattice. Specifically, side chain density for E122-S125 was not observed, and main chain electron density for K124S125 was not sufficient to allow a continuous model to be traced through this region (Figure 6).

Lattice-Packing Interactions Given the novel orientation of ␣4 relative to other regulatory domains and a relatively small interface between the regulatory and effector domains, it is important to note the extent of lattice contacts in these regions. Overall lattice contacts were determined using Areaimol from Figure 4. Comparison of ␣4 Regions of DrrD and PhoB Regulatory Domains Ribbon representations of the regulatory domains of DrrD (left, orange) and PhoB (right, cyan). The ␣4 region and the neighboring residues in the primary sequence are shaded gray in both ribbon diagrams and the corresponding sequence indicated explicitly below the structures for DrrD and PhoB. An open box surrounds the helical region in the sequences. Residues through this sequence, for which DrrD and PhoB occupy different regions in the Ramachandran plot and therefore are responsible for differing orientations of this segment, are circled and/or color coded in the ribbon diagrams, in the sequence, and in the traced Ramachandran plots at the bottom. These colored residues are labeled in the traced Ramachandran plots together with the first and last residues in the sequences. Side chains mediating hydrophobic interactions between ␣4 and the protein core are shown in ball and stick represention. Also shown in ball and stick is DrrD V90, which extends away from the protein core to form crystal lattice contacts (see Figure 7).

Full-Length OmpR Homolog from T. maritima 159

rounding the interdomain interface of a lattice partner, as shown in Figure 7. The regions interacting with this protruding ␣4 of one protomer include the ␤2-␤3 loop of the DNA binding domain and the ␣1-␤2 loop, the N-terminal residues of ␤1, and a single residue (R115) of ␣5 from the regulatory domain of the second protomer. The protected surface area in the vicinity of V90 (␣4 region) of one interacting protomer arising from the lattice interaction shown in Figure 7 amounts to 572 A˚2, which can be compared with the smaller interaction surface observed between domains of a single protein (245 A˚2). Discussion

Figure 5. Effector Domains of DrrD, OmpRC, and PhoBC DrrD effector domain (orange) was superposed using STAMPS [25] with effector domains from OmpR [3] (gray) and PhoB [5] (blue) solved as crystal and NMR structures, respectively, of the isolated domains. A gold, filled cylinder highlights the recognition helix, known to mediate specific interactions with the DNA major groove for this subfamily of transcription regulators. A six-residue loop between H2 and H3 of DrrD that is disordered in the crystal lattice is indicated by a dashed line.

the CCP4 program suite [39] and are displayed on a per residue basis for the entire sequence of DrrD in Figure 2. Of note is a lattice interface involving surface residues of ␣4, centered on V90, and an extensive region sur-

Intradomain Conformational Flexibility The novel orientation of ␣4 in the DrrD regulatory domain contrasts with orientations observed in previously solved regulatory domains; this may result from inherent conformational plasticity of the DrrD ␣4 segment. As noted in Results, the different orientations of the ␣4 segment in DrrD relative to PhoBN can be explained by a two-residue shift in the position within ␣4 that serves as an anchor to the hydrophobic pocket (Figures 2 and 4). Interestingly, the PhoBN hydrophobic anchor, V92, is highly conserved as hydrophobic within a multiple sequence alignment for the OmpR/PhoB subfamily (261 of 269 sequences in the alignment have V, L, M, or I in this position). However, despite the presence of a hydrophobic residue in this position for DrrD (V90), an alternate hydrophobic residue, Y88, anchors ␣4 to the protein core (Figure 4). As the anchor is moved by two residues in the direction of the N terminus for DrrD relative to PhoBN, this end of the ␣4 helix effectively rotates inward toward the protein core, accommodated by changes in the backbone torsion angles surrounding ␣4 and a shortening of the ␤4 strand by one residue in DrrD relative to PhoBN (Figure 2). This leaves DrrD V90 on the surface of the helix (Figure 4), extending away from

Figure 6. Stereo Ribbon Representation of the Interdomain Interface The side chains interacting in the interdomain interface are rendered in ball and stick (carbon is in black, oxygen is in red, nitrogen is in blue, and sulfur is in yellow), including those extending from ␣5 of the regulatory domain and from the conserved ␤ sheet platform (S1S4) at the N terminus of the effector domain. Electron density maps, generated using 2Fo ⫺ Fc Fourier coefficients and contoured at 1.0 ␴, were rendered around the indicated side chains (2.0 A˚ tolerance) and with 3.0 A˚ tolerance around a modeled main chain sequence that spans the interdomain linker region for which limited electron density did not allow complete model building.

Structure 160

Figure 7. Lattice Contacts Involving DrrD ␣4 and the Interface Region Two lattice mates, red and gray, are shown in ribbon representations. Side chains are indicated in ball and stick for residues for which more than 10 A˚2 of surface area was protected from exposure to solvent as a result of lattice contacts between the represented protomers. The P21 unit cell dimensions are indicated by a box.

the protein interior, and is found to provide the hydrophobic locus of an extensive lattice interface (see Figure 7 and histogram of lattice interactions in Figure 2). The atypical orientation of ␣4 raises several questions. First, is this a major or minor conformer for the DrrD ␣4 segment in solution, given the possible stabilization provided by lattice contacts? Second, is a PhoBN-type orientation of ␣4, in which the conserved hydrophobic residue (DrrD V90) serves as the anchor to the hydrophobic pocket, also an energetically accessible conformation for DrrD? Third, is the “atypical” orientation of ␣4 observed in DrrD expected to be widespread for OmpR/ PhoB subfamily members? The current study cannot address the first two questions, which await structural studies of additional crystal forms or solution studies of this protein. The third question, whether this alternative orientation of ␣4 is common within the subfamily and stabilized by a residue corresponding to Y88 in DrrD, is illuminated by an inspection of a sequence alignment for the OmpR/PhoB subfamily. A 269 member multiple sequence alignment of this subfamily reveals that only three members (including DrrD) have hydrophobic or aromatic residues at the position corresponding to Y88 in DrrD (two with Y and one with I), while 253 have D, E, or N. The remainder are polar (T, Q, and S), alanine (one occurrence), or lysine (one occurrence). Thus, this alternative orientation of ␣4 is probably uncommon within the OmpR/PhoB subfamily. If the PhoBN-type ␣4 orientation is also energetically accessible to DrrD, the alternate conformation observed here would underscore the ability of ␣4 to swing out and away from the protein core and may be relevant

to recent insights provided by structures of activated regulatory domains, in which significant movement of ␣4 occurs in conjunction with repositioning of the conserved aromatic on ␤5 [27, 29–32]. This novel orientation of ␣4 may represent an otherwise minor conformer stabilized here by lattice interactions or it may truly represent the major, but atypical, orientation of ␣4 for DrrD, promoted by a nonconserved hydrophobic residue in ␣4. A second region that exhibits conformational mobility within a domain is the H2-H3 loop in the DNA binding domain, where residues 182–187 are disordered in the lattice (Figure 5). This underscores the conformational flexibility in this region as noted in proteolysis studies of OmpR [40], in the previously solved crystal structures of OmpRC [3, 4], and in the solution structure of PhoBC [5]. This loop region has been implicated in interactions with RNA polymerase for both OmpR and PhoB, but with different regions of the polymerase: the ␴70 subunit for PhoB [41] and the C-terminal domain of the ␣ subunit of RNA polymerase for OmpR [42]. The disorder observed in this region in DrrD, as implied in the previously solved structures of OmpR/PhoB subfamily DNA binding domains, supports the notion that conformational mobility in the H2-H3 loop is important in enabling proper interactions with RNA polymerase for transcription initiation [38].

Interdomain Interactions Recent X-ray studies of activated RR regulatory domains have provided significant structural insights about their function as inducible switches [27–32] (for reviews, see [33–35]). However, direct structural information is still lacking with respect to how this structural activation is transmitted to an attached effector domain. In the crystal structure of E. coli NarL, the only other full-length DNA binding RR whose structure is known, the regulatory domain is oriented close to the recognition helix of the effector domain, inhibiting its interaction with DNA [6]. The extensive interdomain interface (972 A˚2) stabilizes this inhibitory interdomain orientation. Only one other multidomain RR structure has been solved, that of methylesterase CheB from Salmonella typhimurium [20], an atypical RR involved in chemotaxis as a receptor-modifying enzyme. For CheB, significant inhibitory interactions between the regulatory domain and the active site of the effector domain were also observed, as was a similarly extensive interdomain interface (827 A˚2). In both of these cases, one effect of phosphorylation must be to favor alternate conformational states that relieve interdomain functional inhibition [6, 20]. However, it is expected that phosphorylation-induced activation mechanisms are more complex than simple interdomain repositioning for many, if not most, RR transcription regulators. Gene control by these regulators typically occurs through assembly of RRs as single or multiple homodimers on specific binding sites upstream of target genes. Thus, phosphorylation may influence interdomain interactions within a monomer as well as protein-protein interactions within and between homodimers, but differentiating these alternatives has been difficult to dissect experimentally. For example, for OmpR, it was recently shown that single-residue muta-

Full-Length OmpR Homolog from T. maritima 161

tions in the effector domain [43] as well as DNA binding to the effector domain [44, 45] can significantly alter phosphorylation rates of the regulatory domain; however, this can result from either intramolecular effects (between domains of a single protomer) or from intermolecular effects (protein-protein or protein-DNA interactions). Further, McCleary and coworkers recently carried out elegant truncation and domain swapping experiments in PhoB, which indicate that ␣5 of the regulatory domain promotes a specific inhibitory effect on the effector domain with respect to DNA binding, suggestive of an intramolecular activation mechanism [46]; however, dimerization of PhoB upon phosphorylation also appears to be critical for activation [47]. The limited extent of the interdomain interface for DrrD (245 A˚2; Figure 1) and disordered linker region (Figure 6) suggest a much less strictly enforced interdomain orientation than for NarL and CheB. In a recent analysis of intraprotein, interdomain interface areas for twodomain proteins found in the CATH database [48], it was found that the smallest interdomain interface for monomeric proteins (sample size 46) was 260 A˚2 (average 1193 ⫾ 630 A˚2) and that somewhat smaller interdomain interfaces (average 902 ⫾ 532) could be observed in crystals for two-domain proteins that function and crystallize as complexes (sample size 105). In this latter group, about 10% had interfaces less than 260 A˚2. The conclusion from the Thornton study was that monomeric, two-domain proteins with interfaces less than about 260 A˚2 are refractory to crystallization as a result of inherent interdomain flexibility. The relative difficulty in crystallizing OmpR/PhoB subfamily members is consistent with general interdomain flexibility for this subfamily. For DrrD, which is monomeric in solution based on gel filtration data (data not shown), extensive lattice contacts in the region of the interface (Figure 7) may serve to limit interdomain dynamics and allow effective crystal lattice propagation. It cannot be ruled out that these lattice interactions near the interface stabilize a minor conformer and that the predominant solution conformers exhibit a more extensive interdomain interface. However, the flexibility observed in the linker region of DrrD and OmpR [40], the accessibility of the DrrD interface to SCN from the bulk crystallization medium, and the general difficulty in crystallizing OmpR/PhoB homologs together suggest that a limited interdomain interface and relative interdomain flexibility is the predominant solution state for DrrD and is likely the case for many OmpR/PhoB subfamily members. Relevance to Models for RR-DNA and RR-RR Interactions Complexes of full-length RRs with DNA have not yet yielded to structure determination, but biochemical and genetic data have provided some insights as to how individual OmpR/PhoB subfamily members assemble on DNA. To date, studied subfamily members appear to assemble as homodimers on target binding sites [5, 49–52]. Furthermore, sequence analysis of DNA binding sites reveals weakly conserved tandem symmetry of ⵑ10 bp sites, and recent DNA cleavage experiments confirmed this tandem DNA binding mode for the OmpR DNA binding domain with a target binding site [53].

Interestingly, the interplay between phosphorylation state and aggregation state may be complex within the subfamily. For example, PhoB has been shown to dimerize upon phosphorylation in solution in the absence of DNA [47, 54], whereas OmpR fails to appreciably dimerize after phosphorylation unless a DNA binding target is present [55]. Ultimately, activation by phosphorylation of the regulatory domain is expected to promote both tandem binding to DNA through the DNA binding domain and specific protein-protein interactions to stabilize dimer formation or higher-order interactions where multiple dimers assemble on multiple binding sites [49]. X-ray structures of isolated regulatory domains have demonstrated specific dimer interfaces in the crystal lattice that, in some cases, have been shown to relate to functionally relevant protein-protein dimer interfaces [29] or in which the functional links have not been directly established [7, 32, 36]. Of note, these observed dimer interfaces all involve rotational, rather than translational (tandem), symmetry, i.e., a two-fold rotation about an axis through the interface between paired regulatory domains. The previously described dimeric interfaces for regulatory domains include an interface mediated by ␣4-␤5 in the structure of phosphorylated FixJN [29] and interfaces involving ␣1 and the ␤5-␣5 loop of both PhoBN [36] and the full-length NarL [7]. If tandem arrangement of the DNA binding domain of an OmpR/PhoB homodimer is the predominant, minimal DNA binding module for the subfamily, as suggested by available genetic and biochemical data, a natural question is whether such an arrangement of the effector domain can be stabilized by any of the previously observed regulatory domain dimerization interfaces involving rotational symmetry. To address this question, we used the available models of regulatory domain dimeric interfaces [7, 29, 32, 36] and superposed full-length DrrD onto each through its regulatory domain. For each of these, it was observed that significant structural unraveling of secondary structural elements at the termini of the two domains of DrrD would have to occur for the two DNA binding domains to assume a tandem relative arrangement similar to that modeled from the DNA cleavage experiments [53] while the DrrD regulatory domains paired in a two-fold fashion observed in the dimer models. That is, although the extent of the DrrD interface is small, suggesting interdomain flexibility, the short interdomain linker (five residues) nonetheless constrains the extent of relative interdomain orientations such that, without lengthening the linker region, dimer assembly through rotational-regulatory, tandem-effector interactions would not be possible. This suggests that, for DrrD, phosphorylation-promoted dimer assembly onto target DNA may be promoted through protein-protein interfaces other than the dimeric interfaces observed in previous structural studies of isolated regulatory domains. The extension of these implications to the entire subfamily is problematic, because a wide variation in the length of interdomain linkers has been noted [38]; some of the longer linkers may allow more extensive interdomain repositioning, which would make feasible a rotationalregulatory/translational-effector dimer arrangement, as has been observed for other bacterial transcription regulators, such as AraC [56]. Resolving this question will

Structure 162

require additional structural and biochemical studies to define these protein-protein and protein-DNA interfaces. Biological Implications Extensive biochemical, genetic, and structural studies, spanning the past 15 years, have provided an increasingly detailed picture of the two-component signal transduction pathway that is central to bacterial adaptation and response mechanisms. At one level, the typical components of the pathway appear to be extremely modular: HKs share sensor domains, ATP binding pockets, and phosphotransfer domains and RRs share phosphorylation-activated regulatory domains and DNA binding effector domains. Recently, the ability to crystallize activated and nonactivated regulatory domains has suggested that there are remarkably common features shared in the activation mechanisms of the isolated regulatory domains. However, detailed mechanistic descriptions are increasingly complex as the modular components are assembled to form a unique signaling pathway. For DNA binding RRs, which appear to bind almost exclusively as dimers to target DNA sites, it is still uncertain how the phosphorylation event within the regulatory domain translates to increased DNA binding affinity for the effector domain. A single previously reported structure of an RR transcription regulator indicates that intramolecular repositioning of the domains must be important to relieve inhibitory interactions. In contrast, the present structure of DrrD from another subfamily of RR transcription regulators indicates that a dynamic interdomain relationship exists for this protein. Indeed, the general difficulty in crystallizing multidomain RRs to date may suggest that this property is the rule rather than the exception for this superfamily. Additional biochemical, biophysical, and structural studies will be required to advance further the understanding of the coupling between phosphorylation and activation for a multidomain RR. Experimental Procedures Cloning, Expression, and Purification The DrrD gene was amplified from chromosomal T. maritima DNA using polymerase chain reaction. DNA primers flanking the gene were engineered to provide melting temperatures of approximately 54⬚C with extensions encoding NdeI and BamHI restriction sites at the 5⬘ and 3⬘ termini, respectively. The resulting amplified DNA was digested with NdeI and BamHI restriction enzymes and ligated with T4 DNA ligase into the similarly digested plasmid pJES307 [57]. The resulting plasmid, pDrrD, was amplified in E. coli strain DH5␣, purified with silica columns (Quiagen, Valencia, CA), and transformed into the salt-inducible strain GJ1158 [58]. Protein overexpression at 37⬚C in cells at mid logarithmic growth phase was induced by addition of 5 M NaCl to a final supplemented NaCl concentration of 300 mM and continued for 3 hr. Whole cells were isolated by centrifugation (30 min at 4,000 ⫻ g) and washed with 100 mM sodium phosphate, pH 7.0. (Subsequent steps performed at 4⬚C unless otherwise noted.) Whole-cell pellets, suspended in 50 mM Tris, pH 8.0 (5 mL/g of wet cell pellet), were lysed by sonication, and the resulting suspension was subjected to ultracentrifugation (30 min at 80,000 ⫻ g). Host proteins were cleared by heat denaturation (70⬚C for 20 min) followed by centrifugation (30 min at 5,000 ⫻ g). The remaining thermostable protein in the soluble portion was precipitated by addition of 60% volume of a saturated solution of

ammonium sulfate. The ammonium sulfate pellet was dissolved in 50 mM Tris, pH 8.0, and purified on a 5 ml Hi Trap Blue dye affinity column (Amersham Pharmacia Biotech, Piscataway, NJ) using a 100 ml gradient of 0–2 M NaCl in Tris, pH 8.0, with 2 mM ␤-mercaptoethanol. The isolated DrrD fractions were concentrated and subjected to gel filtration chromatography using a Superdex75 26/60 column (Amersham Pharmacia Biotech) equilibrated in 12.5 mM Tris, pH 8.0, and 250 mM NaCl. Purified protein was concentrated to 15–20 mg/mL in the same buffer. For production of the selenomethionyl derivative [59], plasmid pDrrD was transformed into methionine auxotroph strain B834(DE3) (Novagen, Madison, WI). The following procedure was adapted from a protocol kindly provided by J. Brannigan, R. Lewis, and A. Wilkinson. Starter cells were generated in standard LB medium, isolated by centrifugation (5 min at 4,000 ⫻ g), and washed with 2⫻-M9 minimal medium. The cells were resuspended in the same medium supplemented with 2 mM MgSO4, 9 ␮M FeSO4-7H2O, 0.4% (w/v) glucose, 1.0 ␮g/mL riboflavin, 1.0 ␮g/ mL niacinamide, 1.0 ␮g/mL pyridoxine monohydrochloride, 1.0 ␮g/ mL thiamine, and 19 of the natural L-amino acids (no L-Met) at 40 ␮g/mL to achieve an initial OD600 of 0.05. The culture was grown at 37⬚C to mid logarithmic phase, and protein expression was induced by addition of IPTG to 1 mM. Cells were isolated 5–6 hr after induction, and protein was isolated and purified following an identical procedure as described above for the native protein. Crystallization and Structure Determination Initial crystallization conditions were identified using commercial sparse matrix screening kits (Hampton Research, Liguna Niguel, CA). After optimization, crystals of approximately 0.15 ⫻ 0.2 ⫻ 0.50 mm were obtained using 100 mM MES, pH 6.5, 10% PEG 3350, and 200 mM KSCN. These were propagated reproducibly with microseeding. Similar crystals of the selenomethionine derivative were obtained under conditions identical to those for the native protein using initial microseeding with native crystals. Multiwavelength anomalous data from selenomethionyl DrrD crystals were collected at beamline X8C at the National Synchrotron Light Source (Brookhaven, NY). Cryopreservation was achieved by transferring crystals briefly (5–20 s) in well solution supplemented with glycerol (final condition: 30% glycerol, 10% PEG 3350, 200 mM KSCN, and 100 mM MES, pH 6.5), and crystals were mounted in 20 ␮m diameter nylon filament loops (0.4–0.5 mm diameter) and transferred to a 100 K nitrogen cryostream. Based on an XAFS scan of the mounted crystal, data were collected at 0.9797, 0.9795, and 0.9640 nm (K edge, peak anomalous, and high-energy remote wavelengths). Selenium atoms in seven of nine methionines were identified using automated MAD phasing program Solve [60] using diffraction data to 2.0 A˚. Statistics reflecting the quality of the data are summarized in Table 1. Initial phase estimates were improved by maximum likelihood solvent-flattening density modification using the program Resolve [61], producing experimental electron density maps of excellent quality. These maps were used for the initial model building using the program O [62], producing a model with Rcryst ⫽ 0.356 (Rfree ⫽ 0.358, test set: 836 reflections). This was followed by a round of torsion angle simulated annealing using CNS 1.0 [63] with the maximum likelihood target function, inclusion of experimental phase information, and an anisotropic bulk solvent correction throughout the refinement. B factors were held constant at 20 A˚2 at this stage. This resulted in improvement of the model to Rcryst ⫽ 0.300 (Rfree ⫽ 0.311). Several rounds of rebuilding followed by simulated annealing were performed. Grouped B factor refinement was then carried out, improving the model to Rcryst ⫽ 0.273 (Rfree ⫽ 0.282). At this point, data extending to 1.5 A˚ were included for refinement (test set: 2026 reflections), and the final stages of refinement were carried out using Refmac5 [64] from the 4.1.1 CCP4 program suite [39]. The maximum likelihood target function was used, and experimental phase information was included through all stages of final refinement. Rigid body refinement using the high-resolution data and coordinates from the model built from 2.0 A˚ data gave Rcryst ⫽ 0.260 (Rfree ⫽ 0.273). This was followed by cycles of restrained refinement and the automatic addition of waters using the integrated ARP routine available through the CCP4 graphical interface (V 1.27), followed by individual scrutiny of added waters. The current model includes 217

Full-Length OmpR Homolog from T. maritima 163

water oxygens. Regions of continuous, bilobal electron density were modeled with thiocyanate ions (six total). Electron density was not observed for all side chain atoms of the following residues: R108, K121, E122, S123, K148, K173, E178, S189, K201, K205, and K209; these atoms were not included in the current model. Additionally, backbone atoms for two regions could not be traced: a two-residue segment at the interdomain linker, K124-S125, and a six-residue segment corresponding to the OmpRC ␣-loop S182-E187. Dual side chain conformers were modeled at E37, selenomethionine (MSe) 39, V49, Mse-79, and V90.

13.

14.

15. Acknowledgments The diffraction data used in the structure determination were collected during the RapiData 2000 Practical Course held at the National Synchrotron Light Source in Brookhaven, NY. We wish to thank the organizer, Robert Sweet, beamline X8C Director, Joel Berendzen, members of the X8C workshop team, and the course instructors, especially Marian Szebenyi, Tom Terwilliger, and Ralf Grosse-Kunstleve, for their invaluable input during the workshop. We thank Vikki Robinson for help throughout this project. Received: September 19, 2001 Revised: November 30, 2001 Accepted: December 2, 2001 References 1. Hoch, J.A., and Silhavy, T.J. (1995). Two-Component Signal Transduction (Washington, DC: American Society for Microbiology Press). 2. Stock, A.M., Robinson, V.L., and Goudreau, P.N. (2000). Twocomponent signal transduction. Annu. Rev. Biochem. 69, 183–215. 3. Martinez-Hackert, E., and Stock, A.M. (1997). The DNA-binding domain of OmpR: crystal structure of a winged-helix transcription factor. Structure 5, 109–124. 4. Kondo, H., Nakagawa, A., Nishihira, J., Nishimura, Y., Mizuno, T., and Tanaka, I. (1997). Escherichia coli positive regulator OmpR has a large loop structure at the putative RNA polymerase interaction site. Nat. Struct. Biol. 4, 28–31. 5. Okamura, H., Hanaoka, S., Nagadoi, A., Makino, K., and Nishimura, Y. (2000). Structural comparison of the PhoB and OmpR DNA-binding/transactivation domains and the arrangement of PhoB molecules on the phosphate box. J. Mol. Biol. 295, 1225– 1236. 6. Baikalov, I., Schroder, I., Kaczor-Grzeskowiak, M., Grzeskowiak, K., Gunsalus, R.P., and Dickerson, R.E. (1996). Structure of the Escherichia coli response regulator NarL. Biochemistry 35, 11053–11061. 7. Baikalov, I., Schroder, I., Kaczor-Grzeskowiak, M., Cascio, D., Gunsalus, R.P., and Dickerson, R.E. (1998). NarL dimerization? Suggestive evidence from a new crystal form. Biochemistry 37, 3665–3676. 8. Pelton, J.G., Kustu, S., and Wemmer, D.E. (1999). Solution structure of the DNA-binding domain of NtrC with three alanine substitutions. J. Mol. Biol. 292, 1095–1110. 9. Mizuno, T., Kaneko, T., and Tabata, S. (1996). Compilation of all genes encoding bacterial two-component signal transducers in the genome of the cyanobacterium, Synechocystis sp. strain PCC 6803. DNA Res. 3, 407–414. 10. Mizuno, T. (1997). Compilation of all genes encoding two-component phosphotransfer signal transducers in the genome of Escherichia coli. DNA Res. 4, 161–168. 11. Lange, R., Wagner, C., de Saizieu, A., Flint, N., Molnos, J., Stieger, M., Caspers, P., Kamber, M., Keck, W., and Amrein, K.E. (1999). Domain organization and molecular characterization of 13 two-component systems identified by genome sequencing of Streptococcus pneumoniae. Gene 237, 223–234. 12. Throup, J.P., Koretke, K.K., Bryant, A.P., Ingraham, K.A., Chalker, A.F., Ge, Y., Marra, A., Wallis, N.G., Brown, J.R., Holmes, D.J., et al. (2000). A genomic analysis of two-compo-

16.

17. 18. 19.

20.

21.

22.

23.

24. 25.

26.

27.

28.

29.

30.

31.

32.

33.

34.

nent signal transduction in Streptococcus pneumoniae. Mol. Microbiol. 35, 566–576. Vidal, O., Longin, R., Prigent-Combaret, C., Hooreman, M., and Lejeune, P. (1998). Isolation of an Escherichia coli K-12 mutant strain able to form biofilms on inert surfaces: involvement of a new ompR allele that increases curli expression. J. Bacteriol. 180, 2442–2449. Dorel, C., Vidal, O., Prigent-Combaret, C., Vallet, I., and Lejeune, P. (1999). Involvement of the Cpx signal transduction pathway of E. coli in biofilm formation. FEMS Microbiol. Lett. 178, 169–175. Huycke, M.M., Sahm, D.F., and Gilmore, M.S. (1998). Multipledrug resistant enterococci: the nature of the problem and an agenda for the future. Emerg. Infect. Dis. 4, 239–249. Novak, R., Henriques, B., Charpentier, E., Normark, S., and Tuomanen, E. (1999). Emergence of vancomycin tolerance in Streptococcus pneumoniae. Nature 399, 590–593. Hulett, F.M. (1996). The signal-transduction network for Pho regulation in Bacillus subtilis. Mol. Microbiol. 19, 933–939. Wanner, B.L. (1993). Gene regulation by phosphate in enteric bacteria. J. Cell. Biochem. 51, 47–54. Gunn, J.S., and Miller, S.I. (1996). PhoP-PhoQ activates transcription of pmrAB, encoding a two-component regulatory system involved in Salmonella typhimurium antimicrobial peptide resistance. J. Bacteriol. 178, 6857–6864. Djordjevic, S., Goudreau, P.N., Xu, Q., Stock, A.M., and West, A.H. (1998). Structural basis for methylesterase CheB regulation by a phosphorylation-activated domain. Proc. Natl. Acad. Sci. USA 95, 1381–1386. Nelson, K.E., Clayton, R.A., Gill, S.R., Gwinn, M.L., Dodson, R.J., Haft, D.H., Hickey, E.K., Peterson, J.D., Nelson, W.C., Ketchum, K.A., et al. (1999). Evidence for lateral gene transfer between Archaea and Bacteria from genome sequence of Thermotoga maritima. Nature 399, 323–329. Lee, P.-J., and Stock, A.M. (1996). Characterization of the genes and proteins of a two-component system from the hyperthermophilic bacterium Thermotoga maritima. J. Bacteriol. 178, 5579–5585. Laskowski, R.A., McArthur, M.W., Moss, D.S., and Thornton, J.M. (1993). PROCHECK: a program to check the stereochemical quality of protein structures. J. Appl. Crystallogr. 26, 282–291. Volz, K. (1993). Structural conservation in the CheY superfamily. Biochemistry 32, 11741–11753. Russell, R.B., and Barton, G.J. (1992). Multiple protein sequence alignment from tertiary structure comparison: assignment of global and residue confidence levels. Proteins 14, 309–323. Volz, K., and Matsumura, P. (1991). Crystal structure of Escherichia coli CheY refined at 1.7 A˚ resolution. J. Biol. Chem. 266, 15511–15519. Lewis, R.J., Brannigan, J.A., Muchova´, K., Bara´k, I., and Wilkinson, A.J. (1999). Phosphorylated aspartate in the structure of a response regulator protein. J. Mol. Biol. 294, 9–15. Kern, D., Volkman, B.F., Luginbuhl, P., Nohaile, M.J., Kustu, S., and Wemmer, D.E. (1999). Structure of a transiently phosphorylated switch in bacterial signal transduction. Nature 40, 894–898. Birck, C., Mourey, L., Gouet, P., Fabry, B., Schumacher, J., Rousseau, P., Kahn, D., and Samama, J.P. (1999). Conformational changes induced by phosphorylation of the FixJ receiver domain. Struct. Fold. Des. 7, 1505–1515. Halkides, C.J., McEvoy, M.M., Casper, E., Matsumura, P., Volz, K., and Dahlquist, F.W. (2000). The 1.9 A˚ resolution crystal structure of phosphono-CheY, an analogue of the active form of the response regulator, CheY. Biochemistry 39, 5280–5286. Cho, H.S., Lee, S.Y., Yan, D., Pan, X., Parkinson, J.S., Justus, S., Wemmer, D.E., and Pelton, J.G. (2000). NMR structure of activated CheY. J. Mol. Biol. 297, 543–551. Lee, S.Y., Cho, H.S., Pelton, J.G., Yan, D., Henderson, R.K., King, D.S., Huang, L., Kustu, S., Berry, E.A., and Wemmer, D.E. (2001). Crystal structure of an activated response regulator bound to its target. Nat. Struct. Biol. 8, 52–56. Robinson, V.L., Buckler, D.R., and Stock, A.M. (2000). A tale of two components: a novel kinase and a regulatory switch. Nat. Struct. Biol. 7, 628–633. Lee, S.Y., Cho, H.S., Pelton, J.G., Yan, D., Berry, E.A., and Wem-

Structure 164

35. 36.

37. 38.

39.

40.

41.

42.

43.

44.

45.

46.

47.

48.

49.

50.

51. 52.

53.

54. 55.

mer, D.E. (2001). Crystal structure of activated CheY. Comparison with other activated receiver domains. J. Biol. Chem. 276, 16425–16431. Johnson, L.N., and Lewis, R.J. (2001). Structural basis for control by phosphorylation. Chem. Rev. 101, 2209–2242. Sola`, M., Gomis-Ru¨th, F.X., Serrano, L., Gonza´lez, A., and Coll, M. (1999). Three-dimensional crystal structure of the transcription factor PhoB receiver domain. J. Mol. Biol. 285, 675–687. Brennan, R.G. (1993). The winged-helix DNA-binding motif: another helix-turn-helix takeoff. Cell 74, 773–776. Martinez-Hackert, E., and Stock, A.M. (1997). Structural relationships in the OmpR family of winged-helix transcription factors. J. Mol. Biol. 269, 301–312. SERC (UK) Collaborative Computational Project, N. (1994). The CCP4 suite: programs for protein crystallography. Acta Crystallogr. D 50, 760–763. Kenney, L.J., Bauer, M.D., and Silhavy, T.J. (1995). Phosphorylation-dependent conformational changes in OmpR, an osmoregulatory DNA-binding protein of Escherichia coli. Proc. Natl. Acad. Sci. USA 92, 8866–8870. Makino, K., Amemura, M., Kim, S.-K., Nakata, A., and Shinagawa, H. (1993). Role of the ␴70 subunit of RNA polymerase in transcription activation by activator protein PhoB in Escherichia coli. Genes Dev. 7, 149–160. Slauch, J.M., Russo, F.D., and Silhavy, T.J. (1991). Suppressor mutations in rpoA suggest that OmpR controls transcription by direct interaction with the alpha subunit of RNA polymerase. J. Bacteriol. 173, 7501–7510. Tran, V.K., Oropeza, R., and Kenney, L.J. (2000). A single amino acid substitution in the C terminus of OmpR alters DNA recognition and phosphorylation. J. Mol. Biol. 299, 1257–1270. Ames, S.K., Frankema, N., and Kenney, L.J. (1999). C-terminal DNA binding stimulates N-terminal phosphorylation of the outer membrane protein regulator OmpR from Escherichia coli. Proc. Natl. Acad. Sci. USA 96, 11792–11797. Qin, L., Yoshida, T., and Inouye, M. (2001). The critical role of DNA in the equilibrium between OmpR and phosphorylated OmpR mediated by EnvZ in Escherichia coli. Proc. Natl. Acad. Sci. USA 98, 908–913. Allen, M.P., Zumbrennen, K.B., and McCleary, W.R. (2001). Genetic evidence that the alpha5 helix of the receiver domain of PhoB is involved in interdomain interactions. J. Bacteriol. 183, 2204–2211. Fiedler, U., and Weiss, V. (1995). A common switch in activation of the response regulators NtrC and PhoB: phosphorylation induces dimerization of the receiver modules. EMBO J. 14, 3696–3705. Jones, S., Marin, A., and Thornton, J.M. (2000). Protein domain interfaces: characterization and comparison with oligomeric protein interfaces. Protein Eng. 13, 77–82. Rampersaud, A., Harlocker, S.L., and Inouye, M. (1994). The OmpR protein of Escherichia coli binds to sites in the ompF promoter region in a hierarchical manner determined by its degree of phosphorylation. J. Biol. Chem. 269, 12559–12566. Holman, T.R., Wu, Z., Wanner, B.L., and Walsh, C.T. (1994). Identification of the DNA-binding site for the phosphorylated VanR protein required for vancomycin resistance in Enterococcus faecium. Biochemistry 33, 4625–4631. Wo¨sten, M.M., and Groisman, E.A. (1999). Molecular characterization of the PmrA regulon. J. Biol. Chem. 274, 27185–27190. Aguirre, A., Lejona, S., Vescovi, E.G., and Soncini, F.C. (2000). Phosphorylated PmrA interacts with the promoter region of ugd in Salmonella enterica serovar typhimurium. J. Bacteriol. 182, 3874–3876. Harrison-McMonagle, P., Denissova, N., Martı´nez-Hackert, E., Ebright, R.H., and Stock, A.M. (1999). Orientation of OmpR monomers within an OmpR:DNA complex determined by DNA affinity cleaving. J. Mol. Biol. 285, 555–566. McCleary, W.R. (1996). The activation of PhoB by acetylphosphate. Mol. Microbiol. 20, 1155–1163. Aiba, H., Nakasai, F., Mizushima, S., and Mizuno, T. (1989). Phosphorylation of a bacterial activator protein, OmpR, by a protein kinase, EnvZ, results in a stimulation of its DNA-binding ability. J. Biochem. 106, 5–7.

56. Soisson, S.M., MacDougall-Shackleton, B., Schleif, R., and Wolberger, C. (1997). Structural basis for ligand-regulated oligomerization of AraC. Science 276, 421–425. 57. Tabor, S., and Richardson, C.C. (1985). A bacteriophage T7 RNA polymerase/promoter system for controlled exclusive expression of specific genes. Proc. Natl. Acad. Sci. USA 84, 1074–1078. 58. Bhandari, P., and Gowrishankar, J. (1997). An Escherichia coli host strain useful for efficient overproduction of cloned gene products with NaCl as the inducer. J. Bacteriol. 179, 4403–4406. 59. Hendrickson, W.A., Horton, J.R., and LeMaster, D.M. (1990). Selenomethionyl proteins produced for analysis by multiwavelength anomalous diffraction (MAD): a vehicle for direct determination of three-dimensional structure. EMBO J. 9, 1665–1672. 60. Terwilliger, T.C., and Berendzen, J. (1999). Automated MAD and MIR structure solution. Acta Crystallogr. D 55, 849–861. 61. Terwilliger, T.C. (2000). Maximum-likelihood density modification. Acta Crystallogr. D 56, 965–972. 62. Jones, T.A., Zou, J.Y., Cowan, S.W., and Kjeldgaard, M. (1991). Improved methods for building protein models in electron density maps and the location of errors in these models. Acta Crystallogr. A 47, 110–119. 63. Bru¨nger, A.T., Adams, P.D., Clore, G.M., DeLano, W.L., Gros, P., Grosse-Kunstleve, R.W., Jiang, J.S., Kuszewski, J., Nilges, M., Pannu, N.S., et al. (1998). Crystallography & NMR system: a new software suite for macromolecular structure determination. Acta Crystallogr. D 54, 905–921. 64. Pannu, N.J., Murshudov, G.N., Dodson, E.J., and Read, A. (1998). Incorporation of prior phase information strengthens maximum-likelihood structure refinement. Acta Crystallogr. D D54, 1285–1294. 65. Carson, M. (1997). Ribbons. Methods Enzymol. 277, 493–505. 66. Barton, G.J. (1993). ALSCRIPT a tool to format multiple sequence alignments. Protein Eng. 6, 37–40. 67. Gouet, P., Fabry, B., Guillet, V., Birck, C., Mourey, L., Kahn, D., and Samama, J.P. (1999). Structural transitions in the FixJ receiver domain. Structure 7, 1517–1526. 68. Kabsch, W., and Sander, C. (1983). Dictionary of protein secondary structure: pattern recognition of hydrogen-bonded and geometrical features. Biopolymers 22, 2577–2637. 69. Diederichs, K., and Karplus, P.A. (1997). Improved R-factors for diffraction data analysis in macromolecular crystallography. Nat. Struct. Biol. 4, 269–275. Accession Numbers Coordinates for selenomethionyl-DrrD and structure factors used for refinement have been deposited in the Protein Data Bank (accession code 1KGS).