HomeBioSource International
HomeHelp[Feedback][For Subscribers][Archive][Search]
Abstract of this Article
Reprint (PDF) Version of this Article
Similar articles found in: 
JBC Online
PubMed
PubMed Citation
This Article has been cited by: 
other online articles
Search Medline for articles by:
Davies, D. R. || Rayment, I.
Alert me when: 
new articles cite this article
Download to Citation Manager

J Biol Chem, Vol. 274, Issue 17, 11904-11913, April 23, 1999


The Three-dimensional Structure of a Tn5 Transposase-related Protein Determined to 2.9-? Resolution*

Douglas R. DaviesDagger¡±, Lisa Mahnke Braam?parallel, William S. Reznikoff?**, and Ivan RaymentDaggerDaggerDagger

From the Dagger Institute for Enzyme Research and Department of Biochemistry, University of Wisconsin, Madison, Wisconsin 53705 and the ? Department of Biochemistry, University of Wisconsin, Madison, Wisconsin 53706

  ABSTRACT
TOP
ABSTRACT
INTRODUCTION
EXPERIMENTAL PROCEDURES
RESULTS
DISCUSSION
REFERENCES

Transposon Tn5 employs a unique means of self-regulation by expressing a truncated version of the transposase enzyme that acts as an inhibitor. The inhibitor protein differs from the full-length transposase only by the absence of the first 55 N-terminal amino acid residues. It contains the catalytic active site of transposase and a C-terminal domain involved in protein-protein interactions. The three-dimensional structure of Tn5 inhibitor determined to 2.9-? resolution is reported here. A portion of the protein fold of the catalytic core domain is similar to the folds of human immunodeficiency virus-1 integrase, avian sarcoma virus integrase, and bacteriophage Mu transposase. The Tn5 inhibitor contains an insertion that extends the beta-sheet of the catalytic core from 5 to 9 strands. All three of the conserved residues that make up the "DDE" motif of the active site are visible in the structure. An arginine residue that is strictly conserved among the IS4 family of bacterial transposases is present at the center of the active site, suggesting a catalytic motif of "DDRE." A novel C-terminal domain forms a dimer interface across a crystallographic 2-fold axis. Although this dimer represents the structure of the inhibited complex, it provides insight into the structure of the synaptic complex.

  INTRODUCTION
TOP
ABSTRACT
INTRODUCTION
EXPERIMENTAL PROCEDURES
RESULTS
DISCUSSION
REFERENCES

Transposition is a process in which a defined DNA sequence, called a transposable element, moves from one location to a second location on the same or another chromosome. Transposable elements occur widely in nature and include the simple insertion sequences or composite transposons of bacteria, certain bacteriophages, transposons, and retrotransposons of eukaryotic cells and retroviruses such as HIV-1.1 Originally described by McClintock (1) in a series of elegant experiments of controlling elements in maize, transposons have been found in all phyla studied to date, including humans. These mobile genetic elements are likely to have played a role in genome evolution and continue to shuffle antibiotic resistance traits among bacteria today (for a general review, see Ref. 2). In eukaryotic species, transposons are not only numerous but also very promiscuous and are known to cause chromosome mutations. Also, the DNA cleavage reactions involved in immunoglobulin gene rearrangement have been shown to occur via a transposition mechanism (3).

Achieving a molecular and structural understanding of transposition has been a formidable challenge in part because of the complexity of the process. Transposition is initiated by the binding of a transposable element-encoded protein called a transposase to specific DNA sequences located at or near the ends of the element.Next, the DNA-bound transposase oligomerizes to form a synapticnucleoprotein complex. Thereafter cleavage of one or both strandsat the transposon ends occurs where the exact cleavage sites area property of the specific element (4, 5). The initial strand cleavage reaction is believed to occur via nucleophilic attack of an activated water molecule on the phosphodiester bond at the end of each element to leave a 3' OH group. As described below, IS4 family elements, such as Tn5 and Tn10, have a more complexmechanism in which formation and cleavage of a hairpin intermediateleads to 5' end release. In the final step, the 3' OH performsa nucleophilic attack on the target DNA, leading to strand transfer.

It has proved troublesome to study the structural properties of these enzymes since it has been difficult to crystallize a full-length protein for any of the transposases or the integrases due to their poor solubility properties (6, 7). This problem might be attributable to the apparent structural flexibility introduced by the presence of distinct modules responsible for the DNA binding and catalytic activities. As a consequence studies have focused on isolated domains that are responsible for part of the function of the protein. This approach has yielded the three-dimensional structures for the catalytic core domains of Mu transposase, and HIV-1 and avian sarcoma virus (ASV) integrases (8-10). These fragments contain that part of the intact molecule responsible for the 3' strand cleavage and transfer reactions, which are both phosphoryl transfer reactions. This has been demonstrated for the truncated forms of HIV integrase and ASV integrase proteins that have been found to retain the ability to perform a "disintegration" reaction that mimics the reverse of the strand transfer step (11-13). Remarkablythese catalytic domains exhibit a common fold that appears tobe related to a broader class of polynucleotidyltransferases thatincludes RNase H, both from Escherichia coli and HIV-1 reversetranscriptase, and recombination factor RuvC (14-18). This has led to speculation that the catalytic mechanism of the transposase/integrase superfamily may be similar to the exonucleolytic cleavage reaction of E. coli DNA polymerase I (17).

The catalytic core domains of the Mu and HIV-1/ASV transposase/integrase enzymes consist of a central five-stranded mixed parallel and antiparallel beta-sheet sandwiched between fouralpha-helices. This fold brings three essential carboxylate residues, two aspartates and one glutamate, into close proximity at a shallow cleft on one surface of the protein. These acidic residues are common to all transposases and form the "DDE" motif believed to be responsible for coordinating the divalent metal ions necessary for catalysis. In the case of RNase H of HIV-1 and ASV integrase, a pair of divalent cations has been observed, coordinated by the three conserved carboxylates (19, 20). A magnesium ion has also been observed within the active site of HIV-1 integrase (21). Although the structures of the individual core domains have proved to be of immense value for understanding this family of proteins, the relationship between the functional segments is lost by the strategy of divide-and-conquer. For example, these structures do not provide information about the possible locations of the DNA binding domains, nor do they show how different domains interact with one another. Thus, to understand transposition in more complete detail, we have undertaken a multidomain structural study of Tn5 transposase.

Besides providing a broader context for understanding transposition in general, structural information about Tn5 transposase has the potential to provide specific understanding of the IS4 family. Two representative IS4 family transposases, those encoded by Tn5 and Tn10, have been the object of extensive genetic studies(for reviews see Refs. 22 and 23). The literature on these elements provides a detailed knowledge base by which to interpret the structure of Tn5 and will allow this structure to serve as a basis for future structure/function analyses. Primary sequence examination of the IS4 transposase family suggests that, although they undoubtedly contain DDE residues functioning in divalent metal coordination, the locations of these residues are placed differently in the primary sequence than those found in retroviral integrases or MuA. In addition, comparison of IS4 transposase primary sequences and genetic studies with Tn10 transposase (24, 25) and Tn5 transposase2 suggests that the IS4 transposases contain some critical motifs (such as the Y(2)R(3)E(6)K motif discussed below) not found in other transposases. Finally, IS4 transposases catalyze two additional phosphoryl transfer reactions, in comparison with retroviral integrases and MuA transposase, to generate blunt-ended transposon DNA as opposed to only nicking the DNA. In these IS4 elements the 3' OH group formed by the initial strand cleavage reaction attacks the complementary strand to cleave the element from the donor DNA leaving a hairpin intermediate (27).3 Presumably, this hairpin intermediate is cleaved by the attack of a second water molecule to expose the 3' OH group and leave a blunt end. The resultant 3' OH acts as a nucleophile in the subsequent end strand transfer reaction by attacking a phosphodiester bond on the target DNA. As is the case with the reactions of retroviralintegrases and Mu transposase, these reactions require only divalentcations as cofactors. Understanding the structure of the Tn5 protein will provide a basis for understanding these unique features of the IS4 family of transposases.

Tn5 is a composite transposable element found in Gram-negative bacteria and consists of two IS50 insertion sequences thatflank, in inverted orientation, three genes encoding antibioticresistances (for reviews of Tn5, see Refs. 22 and 28). Each IS50 is bordered by two related 19-base pair sequences, the outside end (OE) and the inside end. IS50R encodes the 476-amino acid transposase. Purified transposase has been found to be necessary and sufficient for catalysis of Tn5 transposition in vitro in the presence of pairs of OE DNA ends and Mg2+ (29). Transposase releases the transposon from donor DNA leaving blunt ends and inserts it into a 9-base pair staggered cut site in target DNA (30, 31). A closely related transposase, Tn10,is thought to form synaptic complexes in which a monomer is responsiblefor all of the catalytic events at each transposon end (25).In contrast, Mu transposase has been shown to function as a tetramerwith a dimer at each mobile element end (32, 33). Complementationstudies of HIV-1 integrase mutants have suggested that this enzymealso acts as a dimer at each viral end (34).

In E. coli, transposition levels must be tightly regulated in order to prevent excessive chromosome mutagenesis. Tn5 employsa unique means of self-control by expressing a truncated versionof the transposase that functions as an inhibitor. This inhibitorprotein contains 421 amino acid residues and differs from thefull-length transposase only by the absence of the first 55 N-terminalamino acid residues. The inhibitor utilizes a distinct initiationsite relative to the transposase. In vivo, the inhibitor proteinis a natural transdominant negative regulator of transpositionand acts presumably by forming inactive mixed multimers with transposase,not by competitive DNA binding (35). Interestingly, transposaseitself can act as an inhibitor when present at sufficient levels(36, 37).

Many obvious questions remain concerning the molecular basis of Tn5 transposition. In particular what are the protein-protein interactions that occur within the synaptic complex? How are the catalytic centers related to each other in the synaptic complex? What is the structure of the catalytic core? How does the non-productivemultimerization occur? In an effort to answer these questionswe have determined the structure of the intact IS50 Tn5 inhibitor. This protein represents 88% of the full-length transposase sequence. This is the first structure of a naturally expressed biologically active transposase fragment and is the most complete transposase structure known. On the basis of its sequence similarity to other transposases, this protein is predicted to contain all of the critical catalytic core regions of the full-length transposase (24) and has been shown to contain the determinants for dimerization (35, 38). Proteolytic studies also suggest that Tn5 inhibitor has a tertiary structure that is similar to full-length transposase (38).

The inhibitor structure reveals that the catalytic domain of Tn5 transposase shares similar structural features with those of HIV-1, ASV, and Mu transposase/integrase even though they share very low sequence homology, although it does include an additional extended beta-sheet. It also confirms the presence of the DDE catalytic motif of the superfamily and reveals the location of an arginine residue in the active site that is strictly conserved in the IS4 subfamily of transposases. The structure suggests that the catalytic motif is "DDRE" for this group of enzymes. This study extends the common framework for transposition to prokaryotic insertion sequences. The Tn5 inhibitor is dimeric where the interface occurs in the C-terminal region of the protein and is dominated by the interaction of two helices that form a scissor-like interaction. Together, these observations provide insights into catalysis and suggest models for the structural basis of regulation of transposition and for the nucleoprotein architecture within transposition intermediates.

  EXPERIMENTAL PROCEDURES
TOP
ABSTRACT
INTRODUCTION
EXPERIMENTAL PROCEDURES
RESULTS
DISCUSSION
REFERENCES

Protein Purification, Crystallization, and X-ray Data Collection-- The inhibitor protein was prepared and purified as described previously (38, 39). In the final step of the purification, the protein was eluted with a salt gradient from a DEAE-anion exchange column. The pooled fractions containing the inhibitor protein were concentrated to ~16 mg/ml and dialyzed against 100 mM tetraethylammonium sulfate and 20 mM Tris at pH 7.9. The inhibitor protein was crystallized at room temperature by micro batch. Typically 15 ?l of protein at 16 mg/ml was combined with an equal volume of 20% PEG 8000, 100 mM tetraethylammonium sulfate, and 100 mM MES, pH 6.0. Crystals grew spontaneously or were micro-seeded and reached a size of 0.7 ¡Ñ 0.4 ¡Ñ 0.3 mm in 14-28 days. Precession photography determined that the crystals belong to the space group P21212. Unit cell parameters are a = 182.4 ?, b = 72.6 ?, andc = 41.7 ? for native crystals measured at -6 ¢XC, and a = 181.8 ?, b = 71.9 ?, and c = 41.3 ? for the platinum derivative recorded at-160 ¢XC with synchrotron radiation. There is one molecule per asymmetric unit and a solvent content of 57%. Crystals for preparation of heavy atom derivatives and data collection on the laboratory area detector were stabilized in a synthetic mother liquor containing 19% PEG 8000, 100 mM tetraethylammonium sulfate, 300 mM NaCl, and 50 mM MES, pH 6.0. Crystals used for data collection with synchrotron radiation were transferred sequentially into a cryoprotectant solution containing 19% PEG 8000, 100 mM tetraethylammonium sulfate, 300 mM NaCl, 50 mM MES, pH 6.0, and 15% ethylene glycol and flash-cooled to approximately -160 ¢XC in a nitrogen stream (40, 41).

Initial native data and all heavy atom derivative data for MIR phasing were collected to 2.9-3.5-? resolution at -6 ¢XC with a Siemens HiStar area detector at a crystal to detector distance of 18 cm. CuKalpha radiation was generated by a Rigaku RU2000 rotating anode x-ray generator operated at 50 kV and 90 mA and equipped with Siemens G?bel mirrors. Diffraction data frames of width 0.15¢X were recorded for 90-120 s. The frames were processed with XDS (42, 43) and internally scaled with XCALIBRE.4 Tables I-III display the diffraction data sta tistics for the native, heavy atom derivative, and MAD phasing data sets.


View this table:
[in this window]
[in a new window]
Table I
Summary of data collection statistics for MIR data

View this table:
[in this window]
[in a new window]
Table II
Overall data collection statistics for MAD data

View this table:
[in this window]
[in a new window]
Table III
Analysis of the MAD signal

Crystallographic Structure Determination-- A structure of the Tn5 inhibitor protein was initially determined by multiple isomorphous replacement from five heavy atom derivatives and subsequently confirmed by multiple wavelength anomalous dispersion from one heavy atom derivative (45) (Tables I-III). Derivatives were prepared by soaking crystals in a solution of synthetic mother liquor containing one of the following: 0.5 mM MeHgCl, 1 mM Au(CN)2, 1 mM ter(pyridine)PtCl, 0.5 mM di-|mu|-iodobis(ethylenediamine)di-platinum(II) nitrate (PIP), or 1 mM bis(pyridine)PtCl. The heavy atom positions were determined from difference Patterson maps and placed on a common origin with difference Fourier maps. The occupancies and positions of the heavy atom binding sites were refined with the program HEAVY (46). The initial phases were modified by solvent flattening with the algorithm of Kabsch and co-workers (48) and utilized to improve the heavy atom refinement (47,48). Phase calculation statistics for these derivatives are included in Table I. A polyalanine model was built into the subsequentelectron density map with the software package FRODO (49, 50).In the early stages of model building, the heavy atom phases werecombined with model phases with SIGMAA weighting (51). Thereafterthe model was improved through cycles of manual model buildingand least squares refinement with the program TNT (52). The crystallographic R-factor for the model refined against the data collected at -6 ¢XC was 22.1% for all data measured from 30 to 2.9 ?.

In order to confirm the validity of the structure of the Tn5 inhibitor protein, additional independent phasing information was obtained from multiple wavelength anomalous dispersion (MAD) measurements. MAD data were collected from a single crystal soaked in 1 mM ter(pyridine)Pt (II) for 12 h. The x-ray wavelengths were chosen from the x-ray fluorescence spectra of the platinum L-III edge recorded directly from the crystal in order to optimize the anomalous dispersion effects from the platinum atoms. The MAD data were recorded with a 3 ¡Ñ 3 tiled CCD detector on the insertion device on beam-line 19 of the Structural Biology Center at the Advanced Photon Source in Argonne, IL. The crystal to detector distance was 260 mm, and the data were collected with frames of width 1.5¢X. Diffraction data were processed using the HKL 2000 software package (53, 54). The Friedel differences in the reference data set (lambda = 1.0273 ?) were externally local scaled to remove systematic errors. Thereafter the other three data sets were placed on a common scale by local scaling to the reference data set (55). This strategy had a profound effect on the quality of the subsequent electron density map. Phases from the MAD data sets were calculated with the program SOLVE (46, 56) and improved by solvent flattening with the program DM (57, 58). The model of Tn5 inhibitor protein based on the MIR phases was oriented into solvent-flattened map with the program AMORE (59). Visual inspection of the map showed that the tracing of thealpha-carbonbackbone in the initial MIR structure was correct. The electrondensity map was improved by combining MAD phases with model phaseswith SIGMAA weighting (51). A portion of representative electron density is shown in Fig. 1. Thereafter the model was improved through cycles of manual model building and least squares refinement with the program TNT (52). The final structure has a crystallographicR-factor of 19.5% at a resolution of 2.9 ?. Refinement statisticsare listed in Table IV.


View larger version (79K):
[in this window]
[in a new window]
Fig. 1.   Stereo view of the electron density associated with central beta-sheet that forms the base of the predicted divalent cation-binding site. The electron density was calculated with coefficients of the form 2Fo-Fc and displayed with the program MOLDED (69) and MOLSCRIPT (70). 

View this table:
[in this window]
[in a new window]
Table IV
Least squares refinement statistics 
TNT refinement. 

  RESULTS
TOP
ABSTRACT
INTRODUCTION
EXPERIMENTAL PROCEDURES
RESULTS
DISCUSSION
REFERENCES

Overall Structure-- The inhibitor protein contains 421 amino acid residues and corresponds exactly to residues Met56-Ile476 of the full-length transposase. Even though the inhibitor proteinis expressed independently from its own initiation site, and thusis a protein in its own right, the residue numbering utilizedin this paper will be that of the corresponding amino acids inthe Tn5 transposase. The current model for the inhibitor startsat Ser70 and terminates at Gln472. Although much of the structure is well defined, many of the loops exhibit considerable flexibility. This flexibility gives rise to breaks in the electron density between Arg104-Trp124, Val246-Arg256, and Met343-Pro346. In addition to these breaks in the polypeptide chain, the followingamino acids were disordered beyond the beta-carbon: Glu72, Glu88, Asp133, Arg215, Lys216, Lys244, Val246, Gln341, Arg342, Met343, Pro346, Asp347, Asn348, Leu349, Met352, Asp400, and Glu417.

The structure of the inhibitor protein may be divided into two major domains as shown in Fig. 2, a catalytic domain and aC-terminal dimerization domain. Residues Ser70-Gln365 form the catalytic domain. This region is a mixed alpha/beta structure and contains the carboxylate residues that have been implicated in metal binding. The catalytic domain is built from seven alpha-helices and nine strands of mixed parallel-antiparallel beta-sheet. The first five strands of sheet and four of the helices bear striking structural similarity to the HIV-1 integrase, ASV integrase, and Mu transposase cores, as well as to RuvC and RNase H of HIV-1 (also RNase H from E. coli) as discussed below. Residues Arg104 to Trp124 and Leu224 to Leu309 represent insertions relative to the core structures of the other integrases. The first insertion includes a 20-residue disordered loop located between beta1 and beta2. The insertion from Leu224 to Leu309 occurs between beta5 and alpha6 and serves to increase the breadth of the sheet from five to nine strands and to deepen the active site cleft. A longalpha-helix, alpha6, extending from Leu309 to Gly335 lies across the face of the beta-sheet and contributes to the structural foundation of the active site. The hydrogen bonding pattern in this helix is disrupted near the active site between residues 320 and 324. The final secondary structural element in the catalytic domain is helix alpha7, which extends from Glu350 to Ala378. This helix couples the catalytic domain to the C-terminal dimerization domain. There is a prominent bend in this helix at Leu366, and this is taken as the dividing line between the two domains.


View larger version (54K):
[in this window]
[in a new window]
Fig. 2.   Stereo ribbon diagram of Tn5 inhibitor protein. Figs. 2-5 were prepared with the program MOLSCRIPT (70). 

The C-terminal domain (residues Leu366-Gln472) contains five alpha-helices (alpha7 to alpha11) and is responsible for the dimer interface observed in the crystal lattice (Fig. 3). It is an extended domain that conveys the impression that this component of the structure has the potential for flexibility. Helices 9 and 11 form extensive interactions with a neighboring molecule across the crystallographic dyad axis as discussed below.


View larger version (73K):
[in this window]
[in a new window]
Fig. 3.   Ribbon representation of the dimer viewed perpendicular to the 2-fold axis (a) and along the crystallographic 2-fold axis (b). The color scheme is as follows: blue, the structurally conserved catalytic core amino acid residues Ser70-Leu224, Leu309-Gln365; yellowbeta-sheet insertion, Leu224-Leu309; red, C-terminal dimerization domain, Leu366-Gln472. The active site residues, Asp97, Asp188, Arg322, and Glu326, are included in ball-and-stick representation. 

The structure is consistent with results obtained from partial proteolysis of Tn5 transposase and the inhibitor protein where many of the cleavage sites coincide with surface loops. The N-terminalregions of both proteins appear to be susceptible to proteolysiswith proteolytic sites after Arg61 and Lys113 (38). Lys113 coincides with a disordered segment of the inhibitor structure. The major proteolytic cleavage region, residues Lys252-Leu263, corresponds to the flexible loop that contains the disordered residues Val246-Arg256 (38). Likewise, the proteolytic region bounded by residues412-440 is located within the extended C-terminal domain and isrelatively solvent-exposed which accounts for the proteolyticsensitivity. It is noteworthy that the tryptic digestion patternsand cleavage sites of the Tn5 transposase and the inhibitor proteins are very similar which suggests that both proteins contain the same fold.

The Active Site-- Inspection of the Tn5 inhibitor protein structure reveals that three carboxylate residues (Asp97, Asp188, and Glu326) reside in close proximity to one another and are associated with a basic residue, Arg322 (Fig. 4). The three residues map close to the position of the catalytic triad in the ASV integrase structure and correspond to the characteristic DDE motif described for transposases of the IS3 family, for Mu transposase and for the retroelement integrases as well as for the mariner/Tc3 family of eukaryotic transposases (24, 60, 61). Changing Glu326 to alanine results in loss of catalytic activity of Tn5 transposasein vivo.2 Sequence alignment with Tn10 transposase based on an N-terminalregion of homology (38) and a C-terminal extended region of homology called C1 (24) shows that Asp97, Asp188, Glu326, and Arg322 of Tn5 transposase correspond to four conserved residues of Tn10which have been shown to be required for catalytic activity (25).The arginine is strictly conserved throughout the IS4 family (24).Thus the structure of the Tn5 transposase active site confirmsthe presence of the DDE carboxylate cluster and suggests thatthe catalytic motif for the IS4 family should be expanded to DDRE.


View larger version (44K):
[in this window]
[in a new window]
Fig. 4.   Stereo close up view of the active site carboxylate residues and the associated Y(2)R(3)E(6)K motif. The conserved carboxylates, Asp97, Asp188, and Glu326 in Tn5 inhibitor protein are compared with the equivalent residues in the ASV integrase core structure (PDB accession number 1VSD, Ref. 44). The inhibitor is depicted in ribbon and ball-and-stick representation, whereas active site residues for ASV integrase are colored in green

The presence of the arginine side chain prevents the three carboxylate groups from coming as close together as they do in the ASV integrase structure. Unless the side chain of Arg322 undergoes a major conformational change upon binding of divalent metal ion(s) and/or substrate, it is difficult to foresee how the transposase active site could be made to resemble exactly the ASV integrase active site, in terms of its coordination of metal ions. The function of arginine in transposase might be to partially neutralize the negative charge on the acidic residues or to orient the carboxylate groups so that they might support a more open coordination for the divalent cations.

Dimer Interface-- The C-terminal dimerization domain of the Tn5 inhibitor protein observed here has no analog in any of the previously published transposase/integrase structures. This domain contributes to the interface between two molecules across a crystallographic 2-fold axis that is formed by alpha-helices 9 and 11. The long C-terminal helices of adjacent molecules pack against one another from residues Ser458 to Met470 at an angle of 65¢X. Interestingly the C-terminal helices come in very close contact. This is facilitated by the presence of Gly462 at the crossover point which allows for a separation of only 3.9 ? between adjacent alpha-carbons. Helix 9 is nearly perpendicular to the C-terminal helix, and it makes contacts with the C-terminal helix,but not with its counterpart on the symmetry-related molecule. The subunit-subunit interactions are primarily hydrophobic in nature and bury approximately 700 ?2 of solvent-accessible surface area. This modest interaction most likely represents the homodimer interface in the inhibitor protein and may account for the facile interchange between monomers and dimers in solution (62).

  DISCUSSION
TOP
ABSTRACT
INTRODUCTION
EXPERIMENTAL PROCEDURES
RESULTS
DISCUSSION
REFERENCES

Comparison of Transposase/Integrase Catalytic Domains-- One of the most remarkable features of the retroviral integrases and Mu transposase is the observation that, even with very low sequence similarity, a significant degree of secondary and tertiary structure conservation exists between their catalytic domains. Even the functionally divergent proteins RNaseH and RuvC exhibit a similar fold. The common core observed in these integrases and transposases consists of five beta-strands laid out in a threeparallel/three antiparallel configuration sandwiched between fourconserved alpha-helices. This fold forms a shallow groove with the catalytic acidic residues located at its base. The first and fourthbeta-strands contribute the two aspartate residues, and a helix near the C terminus of the catalytic core domain (or coil, in the cases of Mu transposase and HIV-1 integrase structures) contributes the glutamate residue. Given the previously observed structural similarity between these enzymes, it is not surprising that Tn5 transposase inhibitor contains a similar folding motif as shown in Fig. 5. For example the r.m.s. difference between the coordinatesfor 81 structurally equivalent alpha-carbons in Mu transposase and the inhibitor protein is 1.77 ? even though the overall sequence identity for these residues is 9%. A numerical comparison between the Tn5 inhibitor protein and the core structures of retroviral integrases and Mu transposase is given in Table V.


View larger version (52K):
[in this window]
[in a new window]
Fig. 5.   Structural comparison between Tn5 inhibitor protein (a), HIV-1 integrase (b), ASV integrase (c), and Mu transposase proteins (d). The structural features common to all of these proteins are colored in blue. The structures were aligned on the core of the Tn5 inhibitor protein with the program OVRLAP (71). The coordinates for the ASV and HIV-1 integrases, and MU transposase core structures were obtained from the Brookhaven Protein Data Bank (accession numbers 1VSD, 1BIU, and 1ITG, respectively (10, 21, 26, 44)). 

View this table:
[in this window]
[in a new window]
Table V
Structural comparisons between the Tn5 inhibitor protein and the core structures for Mu transposase and the retroviral integrases

There are, however, two insertions that distinguish the Tn5 transposase catalytic domain from the previously reported structures for Mu transposase and HIV-1/ASV integrases. The first of these is a large partially disordered 24-residue loop (Leu101-Trp124) between beta1 and beta2. The corresponding loop varies from two residues in HIV-1 integrase to 15 residues in Mu transposase. In the previous structures, the loop between beta1 and beta2 is ordered in the crystal structure. It is possible that the large disordered region of this loop in the Tn5 transposase only becomes ordered upon binding to DNA. Interestingly, this loop is located near the active site and also near the N terminus of the inhibitor protein. In the full-length protein, such an arrangement positions this loop between the site of DNA cleavage and the presumed location of the N-terminal DNA binding domain. It is therefore conceivable that this loop may help orient the transposon DNA in the active site for catalysis.

There is also an insertion of 86 amino acid residues (Leu224-Leu309), relative to ASV and HIV-1 integrase and Mu transposase, between the conserved fifth beta-strand and the alpha-helix that carries the conserved catalytic glutamic acid residue. This insertion is mostlybeta-strand where the additional residues serve to increase the breadth of the beta-sheet by adding four more antiparallel strands at one edge (Figs. 2 and 3). As a consequence of the curvature of thebeta-sheet, these additional strands wrap around the long alpha-helix,alpha6, that forms the foundation of the active site and forms a distinctwall that overlooks the catalytic carboxylates. These additionalstructural elements change the active site from a shallow depressionobserved in the Mu transposase and retroviral integrase structuresto an elongated canyon in the Tn5 protein. Although the functionof the insertion in Tn5 is unknown, it is interesting that the partially disordered loop (Ile241-Lys260) that lies at the edge of the inserted sheet contains eight positivelycharged residues and suggests that these might contribute to thenonspecific DNA binding component of the transposase. The Mu transposasecontains a traditional beta-barrel in addition to its catalytic domain; however, this is located in a different position. Its subdomain is located at the C terminus of the catalytic domain and is located on the opposite side of the protein relative to the active site such that its function is clearly different from the insertion in the Tn5protein.

The YREK Signature-- The catalytic arginine and glutamate residues discussed above are part of a signature sequence, Y(2)R(3)E(6)K, characteristicof many, but not all, transposases of the IS4 family (24). Mutation of the corresponding Tyr to Phe in Tn10 transposase resulted in a decrease to 83% of wild type transposition activity in vivo (63). In the Tn5 protein structure, Tyr319 is partially buried adjacent to the carboxylate group of Asp188, one of the active site aspartate residues. Since transposition was decreased by only 17% in the tyrosine to phenylalanine mutant of Tn10 transposase, it seems likely that the tyrosine does not play a direct role in catalysis. Interestingly the YS mutant in Tn10 (63) and a YA mutant in Tn52 eliminated the enzymatic activity which suggests that the phenyl group may be important for stabilizing the tertiary structure of the active site. The function of the conserved Lys of the YREK signature is less clear. Mutation of the Lys to Ala in Tn5 transposase resulted in a mutant that impaired cleavage.2 This result is in contrast to a mutation of the Lys to Ala in Tn10 transposase that resulted in a mutant that allowed cleavage but was defective in target capture or strand transfer (25). In the inhibitor protein structure, this residue is solvent-exposed and does not interact with any of the active site residues. Theepsilon amino group is located >10 ? away from the carboxyl group of Asp97 and resides at the base of the active site canyon in the Tn5structure. This amino acid could be involved in retention or orientationof substrate DNA during cleavage or strand transfer.

Possible Interactions between N- and C-terminal Domains-- It is clear that the protein-protein interactions involved in homodimers of the Tn5 transposase and homodimers of the Tn5 inhibitor protein are somewhat different since the Tn5 inhibitor protein can homodimerize in solution under conditions where the transposase is predominantly monomeric (35).5 Yet the proteins differ only by the presence of an additional 55 amino acids in the transposase where these amino acids unambiguouslyparticipate in specific binding to OE DNA (64, 65). This implies that the specific DNA binding domain of the transposase influences dimerization. Inspection of the inhibitor structure shows that the N terminus and C terminus of the inhibitor structure are located near each other, and thus, presumably the N-terminal DNA binding domain and the C-terminal dimerization domain also lie close to one another in transposase. This suggests that the transposase N-terminal and C-terminal domains interact in such a way that the N-terminal domain prevents the C-terminal domain-mediated dimerization. It should also be noted that the C terminus of Tn5 transposase is known to inhibit N terminal-mediated DNA binding (39, 66). The monomeric nature of Tn5 transposase may have functional consequences for transposition. It seems plausible that monomers of transposase bind OE DNA ends and that synapsis of monomer-bound ends leads to productive transposition. Inhibition appears to occur via dead-end complexes through C-terminal heterodimerizationof a monomer-bound end with an inhibitor molecule (67).

Significance of the Dimer Interface-- Protein-protein interactions are important for proper nucleoprotein synaptic complex formation in all transposases. Tn5 transposase is unique in its use of non-productive protein-protein interactions, involving both the inhibitor protein and transposase, to accomplish inhibition and a related phenomenon of transposase cis-restriction in vivo (65). The protein-protein interactions observed in the Tn5 inhibitor structure appear to be involved in the process of inhibition but not synapsis.

The observed dimer conformation of Tn5 inhibitor does not appear to represent a structure that might form the basis for a model of synapsis even though it does contain some attractive elements. For example, inspection of the model shows the catalytic sites are positioned on the same side of the dimer. It would be easy to imagine a concerted strand transfer reaction; however, the distance between the active sites in the dimer is approximately 65 ?, which is too far apart to account for the 9-base pair spacing between the cuts made in the target DNA during strand transfer. If the observed dimerization interface is present at synapsis, then a major domain rearrangement must take place to bring the active sites on the two molecules of the dimer closer together. It is not easy to predict how this might occur, but if the interactionbetween the C-terminal domains is preserved at synapsis, a simpleway to accomplish this might be to rotate the domains downwardand allow the catalytic domains to approach more closely.

A plausible hypothesis is that the dimer interaction in the inhibitor structure represents the structure of the inhibited complex. The role of the C-terminal domain in inhibition is suggested by a point mutation located within the long helix of the C-terminal dimerization motif that was designed on the basis of the structure reported here. The mutant AD466 in the inhibitor protein is observed to prevent homodimerization of the inhibitor protein and eliminates its inhibitory effect on transposition by presumably preventing the formation of the transposase-inhibitor complex on DNA (67).Another line of evidence that implicates the observed dimer interfacein inhibition is a primary sequence alignment analysis of Tn10and Tn5 transposases that indicates that the least conserved regions occur at the C termini. It is of interest that Tn10 transposase is not negatively regulated by protein dimerization but does undergosynapsis. Therefore, the dimer interface observed in Tn5 inhibitorprotein may have no counterpart in Tn10 (38). Thus a role for the C-terminal dimer interface in inhibition but not synapsis is suggested.

Constraints on the Structure of the Synaptic Complex-- Since the protein-protein interactions in the structure are unlikely to be representative of the synaptic complex, it is possible that formation of the synaptic complex would involve a different dimer interaction. Two regions of transposase have been identified as containing determinants for dimerization based on far Western studies of proteolytic products, residues Leu114-Arg314 and residues Thr441-Ile476 (38). Clearly the latter region falls within the dimerization domain observed here. The implication of residues Leu114-Arg314 in the dimerization by the proteolytic studies must be viewed with caution since it is uncertain whether the isolated fragments that implicate this region could fold into functional domains; however, DNA binding studies have shown that the first 387 amino acids of transposase are sufficient for dimerization (66, 68).Although the inhibitor does not bind OE DNA specifically, it interactswith an OE-bound transposase monomer in a ternary complex as shownin gel shift experiments (35, 66). Interestingly completeremoval of the dimerization domain from the transposase by truncationat residue 369 eliminates dimerization of the DNA-protein complex,whereas truncation at 387 retains the ability for dimerization(66, 68). These results suggest that there exists a seconddimerization region. Thus distinct dimerization regions couldbe used for inhibition and synapsis.

Since the fold of the catalytic domain of Tn5 inhibitor protein is similar to those of the HIV-1 and ASV integrase core domains, it is appropriate to consider whether the Tn5 protein might dimerizein a similar manner to those proteins. Dimer interactions observedin the crystal lattices of HIV-1 and ASV involve interactionsof integrase helices alpha1 and alpha5 (8, 9). The corresponding structural elements of Tn5 are helices alpha2 and alpha7. These two elements encompass amino acid residues that fall in the range of 114-314. Although this is an attractive proposal, it is highly unlikely that this arrangement is observed at synapsis since it would place the two active sites too far apart to participate in target capture and strand transfer at points that are only 9 base pairs apart as discussed below. Furthermore, the disposition of helix alpha1 in the inhibitor protein would be inconsistent with dimerizationin this way, because alpha1 is located between alpha2 and alpha7 and would block the interaction between two molecules across this interface. This analysis is complicated by the fact that a synaptic complex of Tn5 transposase is likely to be dimeric, whereas a synapticcomplex of HIV-1 integrase is likely to be tetrameric. Due tothese different stoichiometries, the protein-protein interactionsin integrase and Tn5 transposase synapses may be completelydissimilar.

The recent observation that the strand cleavage reaction proceeds through a hairpin intermediate also places constraints on the arrangement of the catalytic domains at synapsis and strand transfer (27).3 The presence of a hairpin intermediate explains how a single active site can cut two strands of DNA. The initial cleavage presumably occurs via attack of a water molecule, on the first strand of DNA to leave a 3' OH group. Thereafter the resultant 3' OH attacks the complementary strand to form a hairpin that is subsequently cleaved by the attack of a second water molecule to expose the 3' OH group. It seems seems likely that each of these phosphoryl transfer reactions utilizes, in whole or in part, the same constellation of metal ions and protein ligands in an enzymatically similar manner. In all probability, the same active site components are responsible for activation of the 3' OH group in the strand transfer reaction as in the cleavage reaction and implies that the nucleophilic 3' OH group will be bound close to the base of the canyon that is proposed to enclose the OE DNA. Since the strand transfer reactionsoccur at sites 9 base pairs apart on opposite strands, this impliesthat the synaptic complex delivers the two attacking 3' OH groupson approximately opposite sides of the DNA helix, depending onthe structure of the intervening bases. Concerted strand cleavagerequires that both 3' OH groups approach the target DNA at thesame time. Interestingly, the arrangement of the active site canyonsobserved in the inhibitor dimer complex precludes such an attacksince they lie approximately perpendicular to the 2-fold axisof the dimer (Fig. 3b). This disposition of the active sites wouldnot be able to deliver the 3' OH groups to an undistorted sectionof target DNA because the OE DNA would block access to the targetDNA. It is predicted that the catalytic core of the transposasemust be reoriented relative to that observed in the inhibitordimer complex to allow direct approach of the active sites tothe target DNA. A schematic drawing describing some of these ideasis shown in Fig. 6.


View larger version (25K):
[in this window]
[in a new window]
Fig. 6.   Schematic model for the mechanisms of inhibition and transposition in the Tn5 system. Transposase is depicted as a three-domain protein consisting of an N-terminal DNA binding domain (green), the catalytic core domain (blue, with a yellow dot showing the relative position of the active site and a groove indicating where DNA binding might occur), and the C-terminal domain (red). The inhibitor protein lacks the N-terminal domain of transposase. The top half of the figure suggests a mechanism for inhibition. The interaction between the C-terminal domains of the inhibitor protein (top left) is preserved in the inhibited complex containing one molecule of inhibitor protein and one molecule of DNA-bound transposase (top right). The bottom half of the figure suggests a mechanism for transposition. Starting from a complex of one molecule of transposase bound to each end of the transposable element, the synaptic complex is suggested to form by dimerization (lower right). This representation is not meant to imply the precise relationship of the transposase subunits in the complex, other than to suggest that the dimer interface in the synaptic complex is different than that observed in the inhibited complex. The remainder of the figure is consistent with earlier models for transposition (5). 

Evidence that interactions between the catalytic domain and the C-terminal dimerization domain are important for transposition is provided by the phenotype of mutations associated with helixalpha7, Glu350-Ala378, which forms the connection between these two domains. Mutation of Leu372 to proline in the transposase results in a hypertransposing phenotype that is highly trans-active (65). The mutation maps to a region, amino acids 369-387, that was postulated to be important for positioning or stabilization of a dimerization domain (38, 65). Since Leu372 is located in the middle of the helix adjacent to another prolineresidue, it is anticipated that introduction of a proline residueat this point will either cause a greater distortion of the helixor alter the relationship between the catalytic and C-terminaldomains.

Conclusions-- The structure of transposase Tn5 inhibitor protein described here answers many of the obvious questions concerning its tertiary structure and the location and disposition of the catalytic residues. There remain, however, many unanswered questions concerning the relationship of this structure to the biological function of the transposase. It is clear that the conformation of the catalytic domain and its relationship to the dimerization interface must be different in the synaptic complex relative to that seen in the Tn5 inhibitor protein since in the latter the active sites are too far apart. It seems highly likely that the interaction of the transposase with the OE DNA increases the binding affinity of the protein toward a second transposase-OE DNA complex and that this interaction induces concerted excision of the transposon. The present structure limits the possibilities for how this can be accomplished. As such the current study provides a stepping stone toward understanding the molecular basis of transposition by the Tn5 transposase.

  ACKNOWLEDGEMENTS

We thank Dr. Hazel Holden for considerable help and insight in tracing and refining the structural model presented here. We also thank Drs. James B. Thoden and Matthew Benning for help in collecting the x-ray data; Dr. Gary Wesenberg for help with the computational aspects of this project; Dr. Cary Bauer for help with the program SOLVE; and Drs. Dona York and Igor Goryshin for helpful discussions. We gratefully acknowledge the help of Drs. Frank Rotella, Norma Duke, and Andrzej Joachimak at the Structural Biology Beamline, Argonne National Laboratory, in collecting the Multiple Wavelength Data. The use of the Argonne National Laboratory Structural Biology Center Beamlines at the Advanced Photon Source was supported by the U. S. Department of Energy, Basic Energy Sciences, Office of Energy Research, Contract W-31-109-Eng-38.
  FOOTNOTES

* This research was supported in part by National Institutes of Health Grants AR35186 (to I. R.) and GM50692 (to W. S. R.).The costs of publication of this article were defrayed in part by the payment of page charges. The article must therefore be hereby marked "advertisement" in accordance with 18 U.S.C. Section 1734 solely to indicate this fact.

The atomic coordinates and structure factors (code 1b7e) have been deposited in the Protein Data Bank, Brookhaven NationalLaboratory, Upton, NY.

¡± Supported by National Institutes of Health Biotechnology Training Grant GM08349.

parallel Predoctoral trainee supported by National Institutes of Health Grant GM07215.

** Recipient of a Vilas AssociatesAward.

DaggerDagger To whom correspondence may be addressed: Institute for Enzyme Research, 1710 University Ave., Madison, WI 53705. Tel.: 608-262-0529; Fax: 608-265-2904; E-mail: ivan@enzyme.wisc.edu.

2 T. Naumann and W. Reznikoff, unpublishedresults.

3 A. Bhasin, I. Goryshin, and W. Reznikoff, unpublished data.

4 G. Wesenberg and I. Rayment, manuscript in preparation.

5 L. M. Braam and W. Reznikoff, unpublishedresults.

  ABBREVIATIONS

The abbreviations used are: HIV-1, human immunodeficiency virus-1; ASV, avian sarcoma virus; MES, 4-morpholineethanesulfonic acid; r.m.s., root mean square; PEG, polyethylene glycol; MAD, multiple wavelength anomalous dispersion; OE, outside end.
  REFERENCES
TOP
ABSTRACT
INTRODUCTION
EXPERIMENTAL PROCEDURES
RESULTS
DISCUSSION
REFERENCES

  1. McClintock, B. (1951) Cold Spring Harbor Symp. Quant Biol.16, 13-47
  2. Berg, D. E. (1989) in Mobile DNA (Berg, D. E., and Howe, M. M., eds), pp. 185-210, American Society for Microbiology, Washington, D. C.
  3. van Gent, D. C., Mizuuchi, K., and Gellert, M. (1996) Science271, 1592-1594[Abstract]
  4. Mizuuchi, K. (1992) J. Biol. Chem.267, 21273-21276[Medline]
  5. Craig, N. L. (1995) Science270, 253-254[Full Text]
  6. Hickman, A. B., Palmer, I., Engelman, A., Craigie, R., and Wingfield, P. (1994) J. Biol. Chem.269, 29279-29287[Abstract]
  7. Jenkins, T. M., Hickman, A. B., Dyda, F., Ghirlando, R., Davies, D. R., and Craigie, R. (1995) Proc. Natl. Acad. Sci. U. S. A.92, 6057-6061[Medline]
  8. Dyda, F., Hickman, A. B., Jenkins, T. M., Engelman, A., Craigie, R., and Davies, D. R. (1994) Science266, 1981-1986[Medline]
  9. Bujacz, G., Jaskolski, M., Alexandratos, J., Wlodawer, A., Merkel, G., Katz, R. A., and Skalka, A. M. (1995) J. Mol. Biol.253, 333-346[Medline]
  10. Rice, P., and Mizuuchi, K. (1995) Cell82, 209-220[Medline]
  11. Bushman, F. D., Engelman, A., Palmer, I., Wingfield, P., and Craigie, R. (1993) Proc. Natl. Acad. Sci. U. S. A.90, 3428-3432[Medline]
  12. Bushman, F. D., and Wang, B. (1994) J. Virol.68, 2215-2223[Abstract]
  13. Kulkosky, J., Katz, R. A., Merkel, G., and Skalka, A. M. (1995) Virology206, 448-456[Medline]
  14. Ariyoshi, M., Vassylyev, D. G., Iwasaki, H., Nakamura, H., Shinagawa, H., and Morikawa, K. (1994) Cell78, 1063-1072[Medline]
  15. Katayanagi, K., Miyagawa, M., Matsushima, M., Ishikawa, M., Kanaya, S., Ikehara, M., Matsuzaki, T., and Morikawa, K. (1990) Nature347, 306-309[Medline]
  16. Yang, W., Hendrickson, W. A., Crouch, R. J., and Satow, Y. (1990) Science249, 1398-1405[Medline]
  17. Yang, W., and Steitz, T. A. (1995) Structure15, 131-134
  18. Rice, P., Craigie, R., and Davies, D. R. (1996) Curr. Opin. Struct. Biol.6, 76-83[Medline]
  19. Davies, J. F., II, Hostomska, Z., Hostomsky, Z., Jordan, S. R., and Matthews, D. A. (1991) Science252, 88-95[Medline]
  20. Bujacz, G., Alexandratos, J., Wlodawer, A., Merkel, G., Andrake, M., Katz, R. A., and Skalka, A. M. (1997) J. Biol. Chem.272, 18161-18168[Abstract/Full Text]
  21. Goldgur, Y., Dyda, F., Hickman, A. B., Jenkins, T. M., Craigie, R., and Davies, D. R. (1998) Proc. Natl. Acad. Sci. U. S. A.95, 9150-9154[Abstract/Full Text]
  22. Reznikoff, W. S. (1993) Annu. Rev. Microbiol.47, 945-963[Medline]
  23. Kleckner, N., Chalmers, R. M., Kwon, D., Sakai, J., and Bolland, S. (1996) Curr. Top. Microbiol. Immunol.204, 49-82[Medline]
  24. Rezsohazy, R., Hallet, B., Delcour, J., and Mahillon, J. (1993) Mol. Microbiol.9, 1283-1295[Medline]
  25. Bolland, S., and Kleckner, N. (1996) Cell84, 223-233[Medline]
  26. Bernstein, F. C., Koetzle, T. F., Williams, G. J. B., Meyer, E. F., Jr., Brice, M. D., Rogers, J. R., Kennard, O., Shimanouchi, T., and Tasumi, M. (1977) J. Mol. Biol.112, 535-542[Medline]
  27. Kennedy, A. K., Guhathakurta, A., Kleckner, N., and Haniford, D. B. (1998) Cell95, 125-134[Medline]
  28. Berg, D. E. (1989) in Mobile DNA (Berg, D. E., and Howe, M. M., eds), pp. 186-210, American Society for Microbiology, Washington, D. C.
  29. Goryshin, I. Y., and Reznikoff, W. S. (1998) J. Biol. Chem.273, 7367-7374[Abstract/Full Text]
  30. Schaller, H. (1979) Cold Spring Harbor Symp. Quant. Biol.43, 401-408[Medline]
  31. Berg, D. E. (1983) Proc. Natl. Acad. Sci. U. S. A.80, 792-796[Medline]
  32. Baker, T. A., Mizuuchi, M., Savilahti, H., and Mizuuchi, K. (1993) Cell74, 723-733[Medline]
  33. Baker, T. A., Kremenstova, E., and Luo, L. (1994) Genes Dev.8, 2416-2428[Medline]
  34. van Gent, D. C., Vink, C., Groeneger, A. A., and Plasterk, R. H. (1993) EMBO J.12, 3261-3267[Medline]
  35. de la Cruz, N. B., Weinreich, M. D., Wiegand, T. W., Krebs, M. P., and Reznikoff, W. S. (1993) J. Bacteriol.175, 6932-6938[Abstract]
  36. DeLong, A., and Syvanen, M. (1991) Proc. Natl. Acad. Sci. U. S. A.88, 6072-6076[Medline]
  37. Wiegand, T. W., and Reznikoff, W. S. (1992) J. Bacteriol.174, 1229-1239[Abstract]
  38. Braam, L. A. M., and Reznikoff, W. S. (1998) J. Biol. Chem.273, 10908-10913[Abstract/Full Text]
  39. Wiegand, T. W., and Reznikoff, W. S. (1994) J. Mol. Biol.235, 486-495[Medline]
  40. Teng, T. Y. (1990) J. Appl. Crystallogr.23, 387-391
  41. Rodgers, D. W. (1997) Methods Enzymol.276, 183-203
  42. Kabsch, W. (1988) J. Appl. Crystallogr.21, 67-71
  43. Kabsch, W. (1988) J. Appl. Crystallogr.21, 916-924
  44. Bujacz, G., Jaskolski, M., Alexandratos, J., Wlodawer, A., Merkel, G., Katz, R. A., and Skalka, A. M. (1996) Structure15, 89-96
  45. Hendrickson, W. A., and Ogata, C. M. (1997) Methods Enzymol.276, 494-523
  46. Terwilliger, T. C., and Eisenberg, D. (1983) Acta Crystallogr. A39, 813-817
  47. Rould, M. A., Perona, J. J., Soll, D., and Steitz, T. A. (1989) Science246, 1135-1142[Medline]
  48. Kabsch, W., Mannherz, H. G., Suck, D., Pai, E. F., and Holmes, K. C. (1990) Nature347, 37-44[Medline]
  49. Jones, T. A. (1978) J. Appl. Crystallogr.11, 268-272
  50. Jones, T. A. (1985) Methods Enzymol.115, 157-171[Medline]
  51. Read, R. J. (1986) Acta Crystallogr. A42, 140-149
  52. Tronrud, D. E., Ten Eyck, L. F., and Matthews, B. W. (1987) Acta Crystallogr. A43, 489-501
  53. Otwinowski, Z. (1986) Crystal Structure Analysis at 1.9? Resolution of the Complex between E. coli trp Repressor and Its Cognate DNA SequencePh.D. thesis, Yale University
  54. Otwinowski, Z., and Minor, W. (1997) Methods Enzymol.276, 307-326
  55. Rypniewski, W. R., Breiter, D. R., Benning, M. M., Wesenberg, G., Oh, B. H., Markley, J. L., Rayment, I., and Holden, H. M. (1991) Biochemistry30, 4126-4131[Medline]
  56. Terwilliger, T. C. (1997) Methods Enzymol.276, 530-537[Medline]
  57. Cowtan, K., and Main, P. (1998) Acta Crystallogr. D54, 487-493[Medline]
  58. CCP4. (1994) Acta Crystallogr. D50, 760-763
  59. Navaza, J. (1993) Acta Crystallogr. D49, 588-591
  60. Polard, P., and Chandler, M. (1995) Mol. Microbiol.15, 13-23[Medline]
  61. Yurieva, O., and Nikiforov, V. (1996) Biochem. Mol. Biol. Int.38, 15-20[Medline]
  62. Miller, S., Lesk, A. M., Janin, J., and Chothia, C. (1987) Nature328, 834-836[Medline]
  63. Haniford, D. B., Chelouche, A. R., and Kleckner, N. (1989) Cell59, 385-394[Medline]
  64. Zhou, M., and Reznikoff, W. S. (1997) J. Mol. Biol.271, 362-373[Medline]
  65. Weinreich, M. D., Gasch, A., and Reznikoff, W. S. (1994) Genes Dev.8, 2363-2374[Medline]
  66. York, D., and Reznikoff, W. S. (1996) Nucleic Acids Res.24, 3790-3796[Medline]
  67. Braam, L. M., Goryshin, I., and Reznikoff, W. S. (1999) J. Biol. Chem.274, 86-92[Abstract/Full Text]
  68. Weinreich, M. D., Mahnke-Braam, L., and Reznikoff, W. S. (1993) J. Mol. Biol.241, 166-177
  69. Fisher, A. J. (1996) MOLDED, University of Wisconsin
  70. Kraulis, P. J. (1991) J. Appl. Crystallogr.24, 946-950
  71. Rossmann, M. G., and Argos, P. (1975) J. Biol. Chem.250, 7525-7532[Medline]

Copyright (C) 1999 by The American Society for Biochemistry and Molecular Biology, Inc.

Abstract of this Article
Reprint (PDF) Version of this Article
Similar articles found in: 
JBC Online
PubMed
PubMed Citation
This Article has been cited by: 
Search Medline for articles by:
Davies, D. R. || Rayment, I.
Alert me when: 
new articles cite this article
Download to Citation Manager

This article has been cited by other articles:


HomeHelp[Feedback][For Subscribers][Archive][Search]