1. Trang chủ
  2. » Luận Văn - Báo Cáo

Báo cáo khoa học: Tet repressor residues indirectly recognizing anhydrotetracycline pptx

9 184 0

Đang tải... (xem toàn văn)

THÔNG TIN TÀI LIỆU

Nội dung

Tet repressor residues indirectly recognizing anhydrotetracycline Peter Schubert*, Klaus Pfleiderer and Wolfgang Hillen Lehrstuhl fu ¨ r Mikrobiologie, Institut fu ¨ r Mikrobiologie, Biochemie und Genetik, Friedrich-Alexander-Universita ¨ t Erlangen, Germany Two tetracycline repressor (TetR) sequence variants sharing 63% identical amino acids were investigated in terms of their recognition specificity for tetracycline and anhydrotetra- cycline. Thermodynamic complex stabilities determined by urea-dependent unfolding reveal that tetracycline stabilizes both variants to a similar extent but that anhydrotetracycline discriminates between them significantly. Isofunctional TetR hybrid proteins of these sequence variants were constructed and their denaturation profiles identified resi- dues 57 and 61 as the complex stability determinant. Association kinetics reveal different recognition of these TetR variants by anhydrotetracycline, but the binding constants indicate similar stabilization. The identified residues connect to an internal water network, which suggests that the discrepancy in the observed thermo- dynamics may be caused by an entropy effect. Exchange of these interacting residues between the two TetR variants appears to influence the flexibility of this water organiza- tion, demonstrating the importance of buried, structural water molecules for ligand recognition and protein func- tion. Therefore, this structural module seems to be a key requisite for the plasticity of the multiple ligand binding protein TetR. Keywords: Tet repressor; gene regulation; protein stability; ligand binding; antibiotic resistance. The biological function of many proteins is triggered and modulated by binding of effector molecules or a variety of extrinsic cofactors that greatly expand the repertoire of cellular processes executed by polypeptides, DNA or small proteins [1]. Therefore, molecular recognition is a funda- mental process in all living systems, regulating processes as diverse as transcription, cell signalling and immunity [2–4]. Recognition mechanisms may be divided into two general categories, named specific and related specificities [5]. For example, mature immunoglobulins (Ig) are highly specific while those in the germline bind a broad range of antigens [6]. This effect is explained by a diversity of conformations for the related specificity germline antibody, a pre-existing subpopulation of antibody isomers based on increased flexibility [7,8]. To understand the thermodynamic and kinetic principles of protein ligand binding in more detail, concepts of energy landscapes and folding funnels were used [9]. Characteriza- tion of binding sites revealed that such regions are usually depressions in the protein surface where a greater average degree of exposure of hydrophobicity groups occurs [10]. However, for a detailed knowledge of protein specificity at the molecular level it is essential to understand the mechanisms of protein–ligand recognition by obtaining information about the structure, energetics and dynamics of the free and complexed species under a variety of condi- tions. We used two sequence variants of the tetracycline repressor (TetR) to investigate ligand recognition of two different tetracycline (tc) derivatives, tc and anhydrotetra- cycline (atc). TetR regulates resistance to the antibiotic tc in Gram-negative bacteria by inducer binding [11]. This system is successfully adapted for regulation of gene expression in different organisms [12]. Based on sequence similarities of isolates from various bacteria TetR variants were grouped into nine classes called A to E, G, H, J and 30 [13]. The proteins share between 38 and 88% sequence identity and are presumably isostructural and isofunctional. Each homodimeric protein consistes of an N-terminal DNA- binding domain connected to a core domain harbouring the dimerization motif and the effector binding site. Repression of gene expression occurs by specific binding to tetO via a helix–turn–helix motif. Binding of the effector molecule leads to a conformational change resulting in the loss of DNA binding and initiation of transcription [11,14]. The crystal structure of TetR(D)[tc–Mg] 2 indicates the position- ing of the inducer inside the protein core and reveals interactions of the drug with both monomers [15,16]. Here we studied the thermodynamic complex stability and binding affinity of the two naturally occuring TetR variants B and D with tc and atc. The sequence identity between these two TetR variants is only 63%, but it includes most residues involved in operator and DNA binding. The sequence identity for the helix–turn–helix domain is 94% and for the residues contacting tc it is 68%. Urea-dependent unfolding yields similar stabilization of the two sequence variants by tc binding. However, atc shows a strong discrimination between TetR(B) and TetR(D) in terms of stabilization. Using TetR(B/D) hybrid proteins we have Correspondence to W. Hillen, Lehrstuhl fu ¨ r Mikrobiologie, Institut fu ¨ r Mikrobiologie, Biochemie und Genetik, Friedrich-Alexander- Universita ¨ t Erlangen-Nu ¨ rnberg, Staudtstraße 5, 91058 Erlangen, Germany. Fax: + 49 91318528082, Tel.: + 49 91318528081, E-mail: whillen@biologie.uni-erlangen.de Abbreviations: TetR, tetracycline repressor; tc, tetracycline; atc, anhydrotetracycline. *Present address: Biomedical Research Centre, University of British Columbia, Vancouver, V6T 1Z3, Canada. (Received 5 February 2004, revised 21 March 2004, accepted 30 March 2004) Eur. J. Biochem. 271, 2144–2152 (2004) Ó FEBS 2004 doi:10.1111/j.1432-1033.2004.04130.x narrowed the determinants for the different atc complex stabilities to two residues localized at one side of the tc-binding pocket, where they are involved in the last step of the proposed induction mechanism of TetR by stabilizing the induced complex by a water zipper [14]. Thermo- dynamic and kinetic investigations reveal that replacement of these residues in TetR(D) to the ones found in TetR(B) reduces complex stability and recognition of atc. This indicates that this water network is important for stability and drug affinity, but not for induction. Experimental procedures Material and general methods Anhydrotetracycline (atc) was purchased from Acros (Geel, Belgium), all other chemicals were from Merck (Darmstadt, Germany), Roth (Karlsruhe, Germany) or Sigma (Mu ¨ n- chen, Germany) at the highest available purity. Enzymes for DNA restriction and modification were from New England Biolabs (Schwalbach, Germany), Boehringer (Mannheim, Germany), Stratagene (Heidelberg, Germany) or Pharma- cia (Freiburg, Germany). Oligonucleotides were from PE Applied Biosystems (Weiterstadt, Germany). Sequencing was carried out according to the protocol provided by Perkin Elmer for cycle sequencing and sequence waas analysed with an ABI PRISM TM 310 Genetic Analyzer (PE Applied Biosystems, Weiterstadt, Germany). Bacterial strains and plasmids All bacterial strains were derived from Escherichia coli K12. Strain DH5a [hsdR17(r K m K + ), recA1, endA1, gyrA96, thi, relA1, supE44, /80 dLacZ(M15, [(lacZYA-argF)U169] was used for general transformation procedures. Strain WH207 (lacX74, gaK2, rpsL, recA13) [17] served as host strain for b-galactosidase assays. The plasmids pWH806 and pWH853(B) [17] and pWH853(D) [18] used in the in vivo assay as well as pWH1950 [19] for overexpression have been described before. Construction of the chimeric tetR genes All tetR variants were constructed by PCR according to the three-primer method [20]. The products of the second PCR reaction were purified and digested with XbaI/MluIor MluI/NcoI and cloned in pWH853 to replace the respective position of tetR. For overexpression these constructs were digested with XbaIandNcoI and cloned into likewise digested pWH1950. DNA of positive candidates was analysed by sequencing of tetR. b-Galactosidase assay Repression and induction efficiencies of the TetR variants were assayed in E.coliWH207ktet50 carrying the respective pWH853 derivatives. The phage ktet50 contains a tetA– lacZ transcriptional fusion [17] integrated as single copy into the WH207 genome. Bacteria were grown at 28 °Cin Luria–Bertani medium supplemented with the appropriate antibiotics. Quantification of induction efficiencies was carried out with 0.2 lgÆmL )1 atc in overnight and log phase cultures. b-Galactosidase activities were determined as described by Miller [21]. Three independent cultures were assayed for each strain and measurements were repeated at least twice. Purification of the TetR variants pWH1950 derivatives of the different constructs were transformed into E.coliRB791. Cells were grown in 3 L of Lurian–Bertani medium at 28 °C in shaking flasks. TetR expression was induced by adding isopropyl thio-b- D - galactoside to a final concentration of 1 m M at D 600 ¼ 0.7–1.0. Cells were pelleted and resuspended in buffer A containing 50 m M NaCl, 2 m M dithiothreitol, 20 m M sodium phosphate buffer pH 6.8 and broken by sonication. TetR variants were purfied by cation exchange chromato- graphy and gel filtration as described [19]. The amounts of the proteins were obtained from the UV absorption at 280 nm [22] and their activities were determined by titration with atc [23]. Fluorescence and CD spectroscopy Fluorescence intensities were measured with a Spex Fluo- rolog 1680 double spectrometer in 1 cm cells at protein concentrations of 1 l M or 5 l M . Excitation was at 280 nm and emission was recorded at the maximum of the difference between the native and the denatured fluores- cence spectrum. The bandwidth for excitation and emission was 2.2 mm. CD measurements were carried out on a Jasco J-715 spectropolarimeter in 0.5 cm cells at protein concen- trations of 5 l M TetR monomer. The TetR[atc–Mg] 2 complex was formed by adding 10 l M atc. Unfolding of the TetR complexes, thermodynamic and kinetic constants We used F-buffer containing 100 m M NaCl, 100 m M Tris/ HClpH7.5,5m M MgCl 2 ,1m M EDTA, 1 m M dithiothre- itol for all spectroscopic measurements. Urea was obtained from ICN Biochemicals (Eschwege, Germany) and urea solutions were prepared each day. Equilibrium denaturation was performed by incubating protein samples overnight at the indicated urea concentration. Renaturation reactions were achieved by incubating the samples overnight at 8 M urea and then diluting them 200-fold with F-buffer. All reactions were performed at 22 °C and all TetR concentra- tions relate to the monomer. Tetracycline or its derivative (xtc) was used in a onefold molar excess over protein. Thermodynamic calculations of the urea-induced denatur- ation of the TetR[tc/atc–Mg] complex variants were performed as described before [24,25] by extending the calculation as it applies to the monomeric (¼ TetR) as follows: 2TetR N þ 2xtc þ 2Mg 2þ ¼½TetRÀxtcÀMg 2 ¼ 2TetR U þ 2xtc þ 2Mg 2þ where N is native and U is unfolded. The left side of the equation shows the association/ dissociation equilibrium and the right side the folding/ Ó FEBS 2004 TetR anhydrotetracycline recognition (Eur. J. Biochem. 271) 2145 unfolding equilibrium. For the unfolding process the equilibium constant (K U )couldbegivenas: K U ¼½U 2 ½Mg 2þ  2 ½xtc 2 =½TetRÀxtcÀMg 2 The equilibrium constant for ligand-free systems is given as: K U ¼ 2P t f 2 U =f N ½24 where P t is the total protein concentration. Mass balance yields: ½Mg 2þ  t ¼½Mg 2þ þ½C N  ½xtc t ¼½xtcþ½C N  where [C N ] is the concentration of the native complex, resulting in: K U ¼ 2P t f U ð½Mg 2þ  t À P t f N Þ 2 ð½xtc t À P t f N Þ 2 =f N where f N and f U are the fraction native or unfolded, respectively. This equation was used to calculate DGinthe different states of the unfolding pathway. Mg 2+ -independent and -dependent atc equilibrium association constants were determined as published [26]. The association rate constants were determined at 28 °C [27] with equimolar concentrations of TetR monomer and atc in F-buffer as mentioned above. All experiments were repeated at least twice. Fig. 1. TetR crystal structure and amino acid composition. (A) Structure of TetR[tc–Mg] 2 . Monomers are shown in blue and red, helices are indicated by numbers in the blue monomer and Tc is shown in green. (B) Alignment of TetR(B) and TetR(D). Conserved residues are shown in reverse type. Tc binding residues (black filled point) and residues involved in coordinating the water zipper (›) are indicated [14,30]. 2146 P. Schubert et al. (Eur. J. Biochem. 271) Ó FEBS 2004 Results Unfolding of TetR complexed with tc or atc We used urea-dependent denaturation as described before [24,25] to determine the stabilities of TetR[atc–Mg] 2 com- plexes in comparison with TetR[tc–Mg] 2 . As shown by the crystal structure of the latter complex binding of tc occurs inside the TetR dimer (Fig. 1A) [15]. The two TetR B and D sequence variants share 63% amino acid identity (Fig. 1B) and fold into the same quaternary structure in the free and tc-complexed forms [15]. Structural changes during unfold- ing of the complexes were observed by fluorescence and CD. As the fluorescence and CD properties of the two proteins in their complexed forms are analogous, we show only the data for the TetR(D)[atc–Mg] 2 complex. The main fluor- escence of the TetR[atc–Mg] 2 complex originates from atc and four naturally occuring trp residues. The emission spectrum excited at 280 nm shows two maxima at 362 nm and 515 nm (Fig. 2A), resulting from trp emission [22] and energy transfer from trp to atc [28], respectively. In the denatured state at 8 M urea the trp emission maximum shifts to at 354 nm and is slightly increased in intensity, but the 515-nm band is absent, indicating the loss of ligand binding (Fig. 2A). Therefore, the intensity change of the 515-nm band was used to follow denaturation. The TetR[atc–Mg] 2 complexexcitedat455nmshowsalsoanemissionmaxi- mum at 515 nm which is also apparently absent at 8 M urea (Fig. 2B). This fluorescence change of atc was also used to monitor complex denaturation to compare with energy transfer. For the respective tc-complexes unfolding was followed by the change of the energy transfer band at 508 nm and the tc fluorescence at 508 nm excited at 370 nm (data not shown). The TetR[atc–Mg] 2 complex shows CD minima at 208 nm and 222 nm (data not shown) reflecting the high content of a-helical structure [15]. The CD spectra in the presence or absence of the inducer are the same in that curve segment. Since both absorption bands are absent at 8 M urea the ellipticity at 222 nm was also used to quantify TetR[atc–Mg] 2 unfolding. Each of the three probes yields identical results when used to observe denaturation, show- ing monophasic, sigmoidal curves indicating the absence of stable folding intermediates (Fig. 2C). This demonstrates that denaturation of the protein fold as observed by CD, and release of the ligand as observed by fluorescence, occur simultaneously. As expected for a bimolecular reaction, the midpoint of the unfolding transition depends on the protein concentration (Table 1). The transition midpoints shift from 6.2 M to about 6.5 M , when the protein concentration is increased fivefold. To analyse the efficiency of the refolding reaction the TetR(D)[atc–Mg] 2 complex was treated with 8 M urea and subsequently diluted 200-fold with urea-free buffer. The fluorescence emission spectrum and CD of these renatured complexes were identical with those obtained from the native form (Fig. 2A,B). We conclude that denaturation under these conditions is a completely reversible, single-step reaction for the TetR[atc– Mg] 2 and the TetR[tc–Mg] 2 complexes. This allows quan- tification using the two-state model [29] in which only folded complexes, unfolded monomers and free ligand exist at equilibrium in significant concentrations. Thermodynamic stabilities of TetR[tc–Mg] 2 and TetR[atc–Mg] 2 complexes Extrapolation of the urea-induced unfolding curves to 0 M urea (Fig. 2C, inset) to calculate the Gibbs free energy of unfolding DG° U (H 2 O) gave the results shown in Table 1. The DG° U (H 2 O) of 123 kJÆmol )1 for TetR(D)[atc–Mg] 2 Fig. 2. Fluorescence probes to determine complex stability. (A) Fluor- escence emission spectra of TetR(D)[atc–Mg] 2 ; native (top curve), re- natured (dotted line below) and denatured (dashed line). (B) Emission spectra of atc; native (top curve), renatured (dotted line, below) and denatured (dashed line). (C) Urea-induced unfolding curve of TetR(- D)[atc–Mg] 2 . Unfolding was followed by different probes (s,energy transfer; d, atc fluorescence; m, CD) and extrapolated for determin- ation of complex stability (inset). Ó FEBS 2004 TetR anhydrotetracycline recognition (Eur. J. Biochem. 271) 2147 represents a stabilization of 69 kJÆmol )1 compared to the free protein [25]. In contrast, the stabilization of TetR(B) by complex formation with [atc–Mg] + is only 26 kJÆmol )1 . Although free TetR(D) is less stable than TetR(B) [25] it is stabilized more by binding of [atc–Mg] + . The complexes of the two sequence variants with [tc–Mg] + are stabilized by 13 kJÆmol )1 and 12 kJÆmol )1 , respectively. This indicates that the TetR(D) sequence variant must undergo more favourable interactions with atc than does TetR(B), despite of the fact that 63% of the amino acids (Fig. 1B) and the folds of the polypeptide chains are identical. These interactions must also be specific for atc as no difference is observed for the tc complexes. Interaction of TetR(B/D) chimera with [atc–Mg] + We used some TetR(B/D) hybrid proteins from previous work [18,25] and constructed five new chimeras based on the assumption that residues contacting tc in the TetR(D)[tc–Mg] 2 structure (Fig. 3) should be involved in binding of atc as well. Residues contacting [tc–Mg] + are marked with open circles in Fig. 1B. The chimeric proteins analysed in this work for binding and affinity to [tc–Mg] + and [atc–Mg] + areshowninFig.4. To determine both their in vivo repression efficiency and inducibility all chimeric genes were cloned into pWH853 [17] and transformed into E.coliWH207ktet50 containing a chromosomal tetA–lacZ-fusion. The b-galactosidase activ- ities at 28 °C were determined in the presence and absence of 0.2 lgÆmL )1 atc. As shown in Table 2 none of the mutants shows a significantly reduced inducibility com- pared to TetR(D). The repression efficiencies are nearly identical to TetR(B), only TetR(D) is a less efficient repressor and is also less inducible as observed before [18]. The TetR(B/D) chimeras were overexpressed and purified to homogenity as described previously [19]. All purified TetR variants show similar spectral properties in the free and complexed forms (data not shown). The unfolding of the free and complexed proteins was carried out by urea- induced denaturation as mentioned above showing identical denaturation pathways (data not shown) indicated by monophasic denaturation curves. The thermodynamic stabilities given as transition midpoints are summarized in Table 2. Tc binding results in a similar stabilization for all sequence variants. For atc only TetR(B/D)51–208 exhibits TetR(D)[atc–Mg] 2 -like stability, all other chimera show TetR(B)[atc–Mg] 2 -like stability. These data indicate that residues between positions 51 and 63 are responsible for the differences in [atc–Mg] + binding. Identifying single residues for atc recognition Three residues of the segment 51–62 are different between TetR(B) and TetR(D) (Fig. 1B). The mutants TetR(B)57/ Table 1. Thermodynamic stability of the TetR(B)[tc/atc–Mg] 2 and TetR(D)[tc/atc–Mg] 2 complexes. Fluorescence of energy transfer a CD b 1 l M 5 l M 5 l M DG° U [kJÆmol )1 ] Urea 1/2 [ M ] DG° U [kJÆmol )1 ] Urea 1/2 [ M ] DG° U [kJÆmol )1 ] Urea 1/2 [ M ] TetR(B)[tc–Mg] 2 83 ± 7 4.7 83 ± 4 5.0 84 ± 4 5.1 TetR(B)[atc–Mg] 2 96 ± 4 5.0 94 ± 3 5.4 93 ± 6 5.5 TetR(D)[tc–Mg] 2 66 ± 3 4.1 67 ± 4 4.3 69 ± 3 4.6 TetR(D)[atc–Mg] 2 123 ± 9 6.2 122 ± 6 6.5 125 ± 4 6.7 a Urea-dependent unfolding was followed by the change of the energy transfer signal at the wavelength with the maximal difference of fluorescence between the native and the denatured forms, at protein concentrations of 1 l M (excitation at 280 nm) and 5 l M (excitation at 295 nm); tc and atc in onefold excess at 28 °C. b The change of CD was observed at 222 nm. Fig. 3. TetR–tc interactions. Binding to the different monomers is shown by blue and green symbols, respectively, and involved water molecules are depicted as red spheres. For comparison the chemical structure of the atc molecule is shown to the right. The two molecules have similar chemical structures, differing only in that the hydroxyl group at position 6 in tc and the neighbouring hydro- gen bond are eliminated in atc, resulting in an aromatic ring C. 2148 P. Schubert et al. (Eur. J. Biochem. 271) Ó FEBS 2004 59/61D and TetR(D)57/59/61B contain replacements of these three amino acids by the respective other residues. The denaturation results shown in Table 2 demonstrate that these three residues determine the stability difference of the TetR–inducer complexes. When they originate from the TetR(B) sequence the atc-mediated stabilization is small, whereas the TetR(D) residues lead to higher stabilization, hence better recognition. We then constructed all possible double and single TetR variants for these positions in the TetR(D) sequence background. They show the same in vivo repression and inducibility as TetR(D) (Table 2). The denaturation data reveal that complex stability is strongly affected when the residue at position 57 is exchanged, but the alterations at positions 57 and 61 are necessary to yield the fully reduced TetR(B)[atc–Mg] 2 -like stability. Thus, they are the determinant for the improved recognition of atc by TetR(D). Association rate constants for inducer binding to the TetR(D/B) variants From the thermodynamic point of view the complex stability should reflect the equilibrium binding constant (K A ) of atc to TetR. Therefore we determined K A for TetR(B), TetR(D) and TetR(D/B) single, double and triple mutants using an improved assay [26] which takes into account the effect that atc binds Mg 2+ independently of TetR to a small extent. The Mg 2+ independent atc equilibrium constants (K T ) of the TetR(D/B) constructs were determined by titration with inducer in the presence of EDTA following complex formation by atc fluorescence emission at 545 nm excited at 455 nm. The K T values were calculated from these titration curves as described [26] and are listed in Table 3. The Mg 2+ -dependent binding constants were obtained from titrations of TetR complexed with atc in the presence of Mg 2+ [26]. The resulting association constants K A arealsolistedinTable3.The value for TetR(D) is identical to that published before [26]. The single exchanges of the residues at positions 57 and 61 show a partial alteration of affinity from the TetR(D) to the TetR(B) sequence variant, whereas the double mutation shows the full reduction. Single and combined alterations involving position 59 show either no or smaller effects. Association rate constant of TetR[atc–Mg] 2 complex formation The different equilibrium binding constants of atc for both TetR sequence variants could be caused by different association or dissociation rate constants, or both. The time-dependent association rate constants of atc for both sequence variants were determined by measuring the increase in the atc fluorescence upon addition of TetR as described [27]. They were fitted using second-order kinetics for a bimolecular reaction and the results are also presented in Table 3 showing a k ass value of 1 · 10 )6 M )1 Æs )1 for TetR(B) and a sevenfold higher rate constant for TetR(D). Since this accounts for the total difference seen in equilib- rium constants the different stability of TetR[atc–Mg] 2 is based only on different association rates and therefore on molecular recognition. Discussion In this study we used urea-induced unfolding to determine the thermodynamic stabilities of two TetR sequence vari- ants B and D complexed with the inducer atc or tc. The change of the spectral probes used show identical results reflecting the coordinated destruction of the tertiary struc- ture, the loss of the ligand binding and the break down of the secondary structure. The monophasic, sigmoidal curves for urea-dependent unfolding allow the use of a two-state model for calculating thermodynamic complex stabilities. The values for the Gibbs free energy DG° U (H 2 O) for the two TetR sequence variants B and D complexed with tc were determined to 83 kJÆmol )1 and 66 kJÆmol )1 , which com- pared to their free forms [25] reveals that tc stabilizes both sequence variants to similar extents of 13 kJÆmol )1 and 12 kJÆmol )1 , respectively. However, atc binding results in complex stabilities of 96 kJÆmol )1 for the B and 123 kJÆ mol )1 for the D variant, which leads to stabilizations of 26 kJÆmol )1 and 69 kJÆmol )1 , respectively. This surprising stabilization difference of atc with the TetR variants B and D is reflected in the respective association constants. The increased affinity for TetR(D) compared to TetR(B) is accounted for by different associ- ation rate constants, thus indicating the identical overall structures of TetR(D)[tc–Mg] 2 and TetR(B)[tc–Mg] 2 [15] that is used to identify the determinant for this difference in Fig. 4. Overview of the chimeric TetR(B/D) constructs. The respective wild-type proteins are shown by filled (TetR(D)) and open (TetR(B)) bars. The hybrid proteins are shown with their designation given on the right. The top panel indicates the location of the TetR residues interacting with tc (Fig. 1B). Ó FEBS 2004 TetR anhydrotetracycline recognition (Eur. J. Biochem. 271) 2149 ligand stabilization. The sequence alignment of B and D shows that the 16 residues that interact with the inducer (Fig. 3) are highly conserved (Fig. 1B). The fact that both sequence variants form identical primary contacts to the inducer supports the hypothesis that the different affinities must be due to an indirect effect. Functional TetR(B/D) hybrid proteins enabled us to narrow the determinant for the different stabilities to the residues 57 and 61, located at the C terminus of the hinge helix a4 connecting the DNA binding domain with the protein core (Fig. 5). Although located close to residue His64 making contact to tc in the crystal structure [15], these residues are too far away to exert a direct influence. The replacement of the solvent- exposed Val by the chemically similar Ile residue at position 57 should cause just a small effect, but this exchange contributes the most to destabilization. This may be explained by a special feature observed in the crystal structure of the TetR(D)–tetO complex [14]. The comparison of the crystal structures in the tc-induced with the DNA-bound forms leads to a proposed induction mechanism [14]. After [tc–Mg] + insertion into the binding tunnel, ring A of the tc molecule is anchored by hydrogen bonds to different residues including His64, Asn82, Phe86 and Gln116 (Fig. 3). His64 is involved in the conforma- tional change of TetR associated with induction [14] and its interaction with tc fixes the C terminus of helix a4. This state is stabilized by a network of cooperative hydrogen bonds including a chain of eight water molecules (Fig. 5) that is not found in the free form of TetR. Val57 participates in this so-called Ôwater zipperÕ [30], representing the only Table 2. In viv o data and thermodynamic stabilities of different TetR variants. Repression Induction a Free +tc b Difference +atc 2 Difference b-gal [%] b-gal [%] Urea 1/2 [ M ] Urea 1/2 [ M ] DUrea 1/2 [ M ] Urea 1/2 [ M ] DUrea 1/2 [ M ] TetR(D) 8.6 ± 1.0 100 ± 1.1 3.7 4.3 0.5 6.3 2.5 TetR(B)51–208D 0.9 ± 0.1 94 ± 1.5 3.7 4.2 0.4 6.1 2.3 TetR(B)63–208D 0.8 ± 0.0 83 ± 1.1 3.5 4.0 0.5 4.7 1.2 TetR(B)75–208D 0.8 ± 0.0 97 ± 1.4 3.4 3.8 0.4 5.0 1.7 TetR(B)84–208D 0.7 ± 0.1 90 ± 1.2 3.3 3.7 0.4 4.8 1.5 TetR(B)92–208D 0.8 ± 0.0 99 ± 2.3 3.5 3.9 0.4 5.1 1.6 TetR(B)110–208D 0.8 ± 0.0 94 ± 4.1 3.7 4.1 0.4 5.2 1.5 TetR(B)123–208D 0.7 ± 0.0 83 ± 2.2 3.9 4.2 0.3 5.0 1.1 TetR(B)168–208D 1.3 ± 0.1 100 ± 2.3 3.9 4.4 0.5 5.1 1.2 TetR(B)179–208D 1.1 ± 0.1 105 ± 2.0 4.0 4.5 0.5 5.2 1.2 TetR(B) 1.2 ± 0.0 96 ± 2.2 4.2 4.7 0.3 5.2 0.9 TetR(B)57,59,61D 1.8 ± 0.0 90 ± 2.7 4.0 n.d. – 6.2 2.2 TetR(D)57,59,61B 7.8 ± 0.5 92 ± 6.1 3.9 n.d. – 5.0 1.1 TetR(D)57,59B 12 ± 0.7 76 ± 1.8 4.1 n.d. – 5.6 1.5 TetR(D)57,61B 8.2 ± 0.4 92 ± 5.7 3.4 n.d. – 5.1 1.7 TetR(D)59,61B 9.7 ± 0.2 89 ± 1.5 3.5 n.d. – 5.8 2.3 TetR(D)57B 7.6 ± 0.2 79 ± 2.0 3.7 n.d. – 5.4 1.7 TetR(D)59B 9.0 ± 1.2 88 ± 6.1 3.7 n.d. – 6.0 2.3 TetR(D)61B 8.4 ± 0.8 83 ± 1.1 3.4 n.d. – 5.7 2.3 a Induction was determined at 0.2 lgÆmL )1 atc. The 100% expression of b-galactosidase corresponds to 10920 ± 1451 units. b Urea 1/2 - values of chimeric TetR variants were calculated in the absence of the inducer atc by the change of the fluorescence at 330 nm or in the presence of atc from the change of the energy transfer signal at 515 nm for tc and 545 nm for atc excited at 280 nm detected at 28 °C. Table 3. Mg 2+ -independent (K T ) and -dependent (K A ) binding constants, association and calculated dissociation rate constants of atc to TetR variants. K T [· 10 7 M )1 ] K A [· 10 11 M )1 ] k ass [· 10 6 M )1 Æs )1 ] k diss a [· 10 )6 s )1 ] TetR(B) 1.95 ± 0.11 2.30 ± 0.21 1.0 ± 0.2 5.0 ± 0.5 TetR(B)57,59,61D 7.10 ± 0.12 7.62 ± 0.24 TetR(D)57,59,61B 2.20 ± 0.17 2. 94 ± 0.17 TetR(D) 10.0 ± 0.31 9.11 ± 0.30 7.0 ± 0.9 7.0 ± 0.4 TetR(D)57,59B 4.90 ± 0.15 4.70 ± 0.16 TetR(D)57,61B 1.95 ± 0.14 2.12 ± 0.18 TetR(D)59,61B 1.25 ± 0.09 6.78 ± 0.22 TetR(D)57B 3.00 ± 0.16 3.34 ± 0.22 TetR(D)59B 7.85 ± 0.20 7.72 ± 0.24 TetR(D)61B 5.45 ± 0.17 4.74 ± 0.19 a Calculated as k diss ¼ k ass /K A . 2150 P. Schubert et al. (Eur. J. Biochem. 271) Ó FEBS 2004 nonconserved residue of the contacting amino acids (in Fig. 1B marked by ›). The assumption that the exchange to Ile could lead to a distortion of the arrangement in this water network connecting the helixes a4 and the loop between helices a6anda7 is in agreement with the strong decreased complex stability of this single mutation alone. Although Ala61 is not directly involved in interacting with the Ôwater zipperÕ the exchange to Asp could sterically influence residue 59 coordinating water W6 due to the larger size and the introduction of a charged residue. From the thermodynamic point of view, this different stabilization of TetR(B) and TetR(D) should be reflected in their equilibrium association constants K A . However, cal- culating DG from the determined binding constants K A by DG ¼ ) RT*lnK A reveals 65.5 kJÆmol )1 for TetR(B) and 68.9 kJÆmol )1 for TetR(D). Taking into account the stability of the free proteins [25] leads to complex stabilities of 140 kJÆmol )1 for TetR(B)[atc–Mg] 2 and 123 kJÆmol )1 TetR(D)[atc–Mg] 2 , respectively. Only the value for TetR(D)[Mg-atc] 2 resembles the result from the denatura- tion experiment. This discrepancy for TetR(B) could be explained by an enthalpy–entropy compensation effect taking into account that amino acid replacements alter both enthalpy and entropy contributions to ligand binding. As shown for the rat intestinal fatty acid-binding protein the changes in molecular interactions may not necessarily correlate with changes in affinity [31,32]. The two identified residues responsible for the different complex stability of TetR(B) and TetR(D) with atc belong to an internal water network which could be partially destroyed by the replace- ment to the respective residues of TetR(B). This fact might lead to the consequence that the local conformational flexibility of the ligand recognition site is increased due to the observation that binding of buried structural water molecules increase flexibility [33]. This increased flexibility leads to an increase of entropy, which is probably not thecaseinTetR(B)duetoaÔdisorderedÕ water network organization. This compensation explains the reduced DG° U (H 2 O) value for TetR(B)[atc–Mg] 2 deduced from urea-induced denaturation. The water zipper could be part of a functional epitope. Taking into account that interactions between biological molecules cannot be reduced to the description of static molecular structures the function of a protein depends also on the distribution and the populations of its conforma- tional states [34]. Such a mechanism provides multiple pathways and allows a single molecular surface to interact with numerous structurally distinct binding part- ners, accommodate mutations through shifts in the dynamic energy landscape and as such is evolutionarily advantageous [9]. Although we are just beginning to understand the properties that makes these consensus binding sites unique, the role of conformational changes induced upon binding at the protein interface has emerged as a factor of key importance. Because this separation/discrepancy is not seen for tc it is most likely that the water zipper plays an important role for stabilization and ligand affinity but not for induction. This arrangement therefore mediates an indirect recognition mode of TetR. This points out that internal bound water molecules increase protein flexibility which is responsible for specificity of ligand binding. These findings contribute to the basic knowledge for drug design necessary to improve specificity of the TetR system. Acknowledgements We thank Prof F. X. Schmid (University Bayreuth) for helpful and stimulating discussions and Dr Oliver Scholz for help in calculating the binding constants. This work was supported by the Deutsche Forschungsgemeinschaft through SFB 473 and the Fonds der Chemischen Industrie. References 1. Zhang, J. & Matthews, C.R. (1998) The role of ligand binding in the kinetic folding mechanism of human p21 (H-ras) protein. Biochemistry 37, 14891–14899. 2. Garvie, C.W. & Wolberger, C. (2001) Recognition of specific DNA sequences. Mol. Cell 8, 937–946. 3. Christophe, D., Christophe-Hobertus, C. & Pichon, B. (2000) Nuclear targeting of proteins: how many different signals? Cell Signal. 12, 337–341. 4. Kleanthous, C. & Walker, D. (2001) Immunity proteins: enzyme inhibitors that avoid the active site. Trends Biochem. Sci. 26, 624–631. 5. Ma, B., Shatsky, M., Wolfson, H.J. & Nussinov, R. (2002) Mul- tiple diverse ligands binding at a single protein site: a matter of pre- existing populations. Protein Sci. 11, 184–197. Fig. 5. Detailed view of TetR(D)[tc–Mg] 2 . The water-network con- necting helices a4anda7 and involved residues V57 (red) and Q109 (orange), and the neighbouring residue A61 (green) are shown. Ó FEBS 2004 TetR anhydrotetracycline recognition (Eur. J. Biochem. 271) 2151 6. Ma, B., Kumar, S., Tsai, C.J. & Nussinov, R. (1999) Folding funnels and binding mechanisms. Protein Eng. 12, 713–720. 7. Foote, J. & Milstein, C. (1994) Conformational isomerism and the diversity of antibodies. Proc. Natl Acad. Sci. USA 91, 10370– 10374. 8. Wedemeyer, W.J., Ashton, R.W. & Scheraga, H.A. (1997) Kinetics of competitive binding with application to thrombin complexes. Anal Biochem. 248, 130–140. 9. Kumar, S., Ma, B., Tsai, C.J., Sinha, N. & Nussinov, R. (2000) Folding and binding cascades: dynamic landscapes and popula- tion shifts. Protein Sci. 9, 10–19. 10. Ringe, D. (1995) Structure-aided drug design: crystallography and computational approaches. J. Nucl. Med. 36, 28S–30S. 11. Hillen, W. & Berens, C. (1994) Mechanisms underlying expression of Tn10 encoded tetracycline resistance. Annu. Rev. Microbiol. 48, 345–369. 12. Berens, C. & Hillen, W. (2003) Gene regulation by tetracyclines. Constraints of resistance regulation in bacteria shape TetR for application in eukaryotes. Eur. J. Biochem. 270, 3109–3121. 13. Levy, S.B., McMurry, L.M., Barbosa, T.M., Burdett, V., Courvalin, P., Hillen, W., Roberts, M.C., Rood, J.I. & Taylor, D.E. (1999) Nomenclature for new tetracycline resistance determinants. Antimicrob. Agents Chemother. 43, 1523–1524. 14. Orth, P., Schnappinger, D., Hillen, W., Saenger, W. & Hinrichs, W. (2000) Structural basis of gene regulation by the tetracycline inducible Tet repressor-operator system. Nat. Struct. Biol. 7, 215–219. 15. Hinrichs, W., Kisker, C., Duvel, M., Muller, A., Tovar, K., Hillen, W. & Saenger, W. (1994) Structure of the Tet repressor-tetra- cycline complex and regulation of antibiotic resistance. Science 264, 418–420. 16. Kisker, C., Hinrichs, W., Tovar, K., Hillen, W. & Saenger, W. (1995) The complex formed between Tet repressor and tetra- cycline-Mg2+ reveals mechanism of antibiotic resistance. J. Mol. Biol. 247, 260–280. 17. Wissmann, A., Wray, L.V. Jr, Somaggio, U., Baumeister, R., Geissendorfer, M. & Hillen, W. (1991) Selection for Tn10 tet repressor binding to tet operator in Escherichia coli:isolationof temperature-sensitive mutants and combinatorial mutagenesis in the DNA binding motif. Genetics 128, 225–232. 18. Schnappinger, D., Schubert, P., Pfleiderer, K. & Hillen, W. (1998) Determinants of protein-protein recognition by four helix bundles: changing the dimerization specificity of Tet repressor. EMBO J. 17, 535–543. 19. Ettner,N.,Muller,G.,Berens,C.,Backes,H.,Schnappinger,D., Schreppel, T., Pfleiderer, K. & Hillen, W. (1996) Fast large-scale purification of tetracycline repressor variants from overproducing Escherichia coli strains. J Chromatogr. A. 742, 95–105. 20. Landt, O., Grunert, H.P. & Hahn, U. (1990) A general method for rapid site-directed mutagenesis using the polymerase chain reac- tion. Gene 96, 125–128. 21. Miller, J.H. (1992) A Short Course in Bacterial Genetics. A Laboratory Manual and Handbook for Escherichia Coli and Related Bacteria. Cold Spring Harbor Laboratory Press, Cold Spring Harbor. 22. Hansen, D., Altschmied, L. & Hillen, W. (1987) Engineered Tet repressor mutants with single tryptophan residues as fluorescent probes. Solvent accessibilities of DNA and inducer binding sites and interaction with tetracycline. J. Biol. Chem. 262, 14030–14035. 23. Takahashi, M., Degenkolb, J. & Hillen, W. (1991) Determination of the equilibrium association constant between Tet repressor and tetracycline at limiting Mg 2+ concentrations: a generally applic- able method for effector-dependent high-affinity complexes. Anal. Biochem. 199, 197–202. 24. Backes, H., Berens, C., Helbl, V., Walter, S., Schmid, F.X. & Hillen, W. (1997) Combinations of the alpha-helix-turn-alpha- helix motif of TetR with respective residues from LacI or 434Cro: DNA recognition, inducer binding, and urea-dependent dena- turation. Biochemistry 36, 5311–5322. 25. Schubert, P., Schnappinger, D., Pfleiderer, K. & Hillen, W. (2001) Identification of a stability determinant on the edge of the Tet repressor four-helix bundle dimerization motif. Biochemistry 40, 3257–3263. 26. Scholz, O., Schubert, P., Kintrup, M. & Hillen, W. (2000) Tet repressor induction without Mg 2+ . Biochemistry 39, 10914–10920. 27. Takahashi, M., Altschmied, L. & Hillen, W. (1986) Kinetic and equilibrium characterization of the Tet repressor-tetracycline complex by fluorescence measurements. Evidence for divalent metal ion requirement and energy transfer. J. Mol. Biol. 187, 341–348. 28. Kintrup, M., Schubert, P., Kunz, M., Chabbert, M., Alberti, P., Bombarda, E., Schneider, S. & Hillen, W. (2000) Trp scanning analysis of Tet repressor reveals conformational changes asso- ciated with operator and anhydrotetracycline binding. Eur. J. Biochem. 267, 821–829. 29. Pace, C.N. (1986) Determination and analysis of urea and guanidine hydrochloride denaturation curves. Methods Enzymol. 131, 266–280. 30. Orth, P., Cordes, F., Schnappinger, D., Hillen, W., Saenger, W. & Hinrichs, W. (1998) Conformational changes of the Tet repressor induced by tetracycline trapping. J. Mol. Biol. 279, 439–447. 31. Richieri, G.V., Low, P.J., Ogata, R.T. & Kleinfeld, A.M. (1997) Mutants of rat intestinal fatty acid-binding protein illustrate the critical role played by enthalpy-entropy compensation in ligand binding. J. Biol. Chem. 272, 16737–16740. 32. Dill, K.A. (1997) Additivity principles in biochemistry. J. Biol. Chem. 272, 701–704. 33. Fischer, S. & Verma, C.S. (1999) Binding of buried structural water increases the flexibility of proteins. Proc. Natl Acad. Sci. USA 96, 9613–9615. 34. Van Regenmortel, M.H. (1999) Molecular recognition in the post- reductionist era. J. Mol. Recognit. 12,1–2. 2152 P. Schubert et al. (Eur. J. Biochem. 271) Ó FEBS 2004 . Tet repressor residues indirectly recognizing anhydrotetracycline Peter Schubert*, Klaus Pfleiderer and Wolfgang Hillen Lehrstuhl. sequence variants of the tetracycline repressor (TetR) to investigate ligand recognition of two different tetracycline (tc) derivatives, tc and anhydrotetra- cycline (atc). TetR regulates resistance. 91318528081, E-mail: whillen@biologie.uni-erlangen.de Abbreviations: TetR, tetracycline repressor; tc, tetracycline; atc, anhydrotetracycline. *Present address: Biomedical Research Centre, University

Ngày đăng: 30/03/2014, 13:20