Overexpression, Purification, Characterization, and Crystallization of

0 downloads 0 Views 621KB Size Report
Jul 10, 1997 - of the BTB/POZ Domain from the PLZF Oncoprotein*. (Received ... development (5–11). Biological ... equilibrium analytical ultracentrifugation, equilibrium dena- .... ance at 280 nm using a molar extinction coefficient of 7450 M 1 cm 1 ... formed by incubating individual samples of the PLZF-BTB/POZ protein.
THE JOURNAL OF BIOLOGICAL CHEMISTRY © 1997 by The American Society for Biochemistry and Molecular Biology, Inc.

Vol. 272, No. 43, Issue of October 24, pp. 27324 –27329, 1997 Printed in U.S.A.

Overexpression, Purification, Characterization, and Crystallization of the BTB/POZ Domain from the PLZF Oncoprotein* (Received for publication, July 10, 1997)

Xinmin Li‡§, Jesus M. Lopez-Guisa‡, Nisha Ninan‡, Evan J. Weiner‡, Frank J. Rauscher, III‡, and Ronen Marmorstein‡§¶i From ‡The Wistar Institute, ¶Departments of Chemistry, and §Biochemistry and Biophysics, University of Pennsylvania, Philadelphia, Pennsylvania 19104

The BTB/POZ domain defines a conserved region of about 120 residues and has been found in over 40 proteins to date. It is located predominantly at the N terminus of Zn-finger DNA-binding proteins, where it may function as a repression domain, and less frequently in actin-binding and poxvirus-encoded proteins, where it may function as a protein-protein interaction interface. A prototypic human BTB/POZ protein, PLZF (promyelocytic leukemia zinc finger) is fused to RARa (retinoic acid receptor a) in a subset of acute promyelocytic leukemias (APLs), where it acts as a potent oncogene. The exact role of the BTB/POZ domain in protein-protein interactions and/or transcriptional regulation is unknown. We have overexpressed, purified, characterized, and crystallized the BTB/POZ domain from PLZF (PLZFBTB/POZ). Gel filtration, dynamic light scattering, and equilibrium sedimentation experiments show that PLZF-BTB/POZ forms a homodimer with a Kd below 200 nM. Differential scanning calorimetry and equilibrium denaturation experiments are consistent with the PLZFBTB/POZ dimer undergoing a two-state unfolding transition with a Tm of 70.4 °C, and a DG of 12.8 6 0.4 kcal/ mol. Circular dichroism shows that the PLZF-BTB/POZ dimer has significant secondary structure including about 45% helix and 20% b-sheet. We have prepared crystals of the PLZF-BTB/POZ that are suitable for a high resolution structure determination using x-ray crystallography. The crystals form in the space group I222 or I212121 with a 5 38.8, b 5 77.7, and c 5 85.3 Å and contain 1 protein subunit per asymmetric unit with approximately 40% solvent. Our data support the hypothesis that the BTB/POZ domain mediates a functionally relevant dimerization function in vivo. The crystal structure of the PLZF-BTB/POZ domain will provide a paradigm for understanding the structural basis underlying BTB/ POZ domain function.

The PLZF BTB/POZ domain, named for its presence in the Drosophila proteins Broad Complex, tramtrack, and bric a brac (BTB) (1) and its homology with several poxvirus proteins and zinc finger proteins (POZ) (2), is an evolutionarily conserved motif of about 120 residues (Fig. 1) found in an increasing number of proteins having a variety of functions (1– 4). Proteins containing a BTB/POZ domain have been identified in poxvirus, Caenorhabditis elegans, Drosophila, and human and * The costs of publication of this article were defrayed in part by the payment of page charges. This article must therefore be hereby marked “advertisement” in accordance with 18 U.S.C. Section 1734 solely to indicate this fact. i To whom correspondence should be addressed: The Wistar Institute, 3601 Spruce St., Philadelphia, PA 19104. Tel.: 215-898-5006; Fax: 215898-3868; E-mail: [email protected].

are generally found at the N terminus of either actin-binding or, more commonly, nuclear DNA-binding proteins (4). Proteins containing a BTB/POZ domain have been associated with diverse functions including nucleosome/chromosome disruption, pattern formation, metamorphosis, oogenesis, and eye and limb development (5–11). Biological relevance for the function of BTB/POZ domains has come from the study of the PLZF (for promyelocytic leukemia zinc finger) oncoprotein, associated with acute promyelocytic leukemia (APL).1 APL results from the malignant proliferation of cells blocked at the stage of promyelocytic differentiation and accounts for about 10% of all myeloid leukemias (12–14). APL is characterized by a non-random reciprocal chromosomal translocation, t(15;17), observed in greater than 95% of patients (15, 16), resulting in the fusion of the retinoic acid receptor a (RARa) gene with a gene called PML (for promyelocytic leukemia). In a small subset of APL cases, a second translocation, t(11;17), has been identified that fuses the RARa gene to a previously uncharacterized gene, PLZF, resulting in the expression of the chimeric PLZF-RARa fusion protein harboring the BTB/POZ domain from the PLZF protein (17). This chimeric protein appears to function as a dominant negative transcription factor (18). Remarkably, APL patients expressing the PLZF-RARa fusion protein are not responsive to treatment with all-trans-retinoic acid, and thus have a poor prognosis for survival (19, 20). The wild-type function of PLZF is not known; however, its homology with the Drosophila gap gene Kruppel (21, 22) and the observation that it functions as a repressor of transcription (23, 24) suggests that it plays a role in developmental programs of gene expression. Recent studies suggest that PLZF may play a specific role in hematopoiesis, perhaps in the maintenance of the phenotype of uncommitted hematopoietic progenitors and/or controlling the commitment of these cells to differentiation (25–27). Moreover, Dong and co-workers have recently shown that the BTB/POZ region of the PLZF-RARa protein display a dominant-negative effect against retinoic acid receptor function (24). These observations, together with the fact that the BTB/POZ domain is present in all APL patients so far identified that express PLZF-RARa fusion proteins (18), strongly suggests that the BTB/POZ domain of the PLZF protein plays an important role in the pathogenesis of APL. We have employed gel filtration, dynamic light scattering, equilibrium analytical ultracentrifugation, equilibrium denaturation, and differential scanning calorimetry measurements to show that the isolated BTB/POZ domain from the human PLZF protein forms a very tight and highly stable homodimeric species. Analysis with circular dichroism also suggests that the

1 The abbreviations used are: APL, acute promyelocytic leukemia; DSC, differential scanning calorimetry.

27324

This paper is available on line at http://www.jbc.org

Characterization of the BTB/POZ Domain from PLZF

27325

FIG. 1. Sequence alignment of the BTB/POZ domains. Among the 401 BTB/POZ domains known to date, a panel of 19 representative sequences that are most homologous with PLZF are aligned (CLUSTAL program) and displayed (BOXSHADE program). Black and gray backgrounds are used to indicate identical and/or conserved residues found in at least 50% of the proteins at a given position, respectively.

BTB/POZ dimer has a high degree of secondary structure including a significant helical content. To facilitate a three-dimensional structure determination of the PLZF BTB/POZ domain, we have prepared crystals of the PLZF BTB/POZ domain that are suitable for x-ray crystallographic analysis. The crystals form in the space group I222 or I212121 with a 5 38.8, b 5 77.7, and c 5 85.3 Å and contain 1 protein subunit per asymmetric unit with approximately 40% solvent. We have collected a complete 2.25 Å data set from a crystal that has been frozen at 2170 °C. EXPERIMENTAL PROCEDURES

Overexpression—The DNA segment encoding residues 1–118 of PLZF comprising the BTB/POZ region (PLZF-BTB/POZ) was amplified from a pSG5 expression vector harboring the human PLZF cDNA (gift from A. Zelent, Institute of Cancer Research) using polymerase chain reaction with pairs of oligonucleotide primers in which PLZF specific primers were linked to BamHI restriction sites. This amplified fragment was subcloned into the BamHI site of the PQE30 T5-polymerasebased expression vector and transformed into Escherichia coli S9 cells (28). Plasmid from single clones were checked for insertion and correct directionality of the PLZF-BTB/POZ insert by restriction digest. The sequence of clones that contained the PLZF-BTB/POZ insert in the correct orientation were confirmed by dideoxynucleotide sequencing (Wistar DNA Core Facility). The expressed protein contained a 12residue N-terminal His6 tag sequence MRGSHHHHHHGS, followed by residues 1–118 of the PLZF protein (PLZF-BTB/POZ). Recombinant PLZF-BTB/POZ protein was produced by growing PLZF-BTB/POZ-transformed S9 cells to an A595 of 0.5 and inducing with 0.5 mM isopropyl-1-thio-b-D-galactopyranoside for 3 h. Small scale inductions in Luria broth followed by cell disruption by sonication and centrifugation to separate the soluble and insoluble protein fraction showed that the protein was partitioned about 50/50 between the soluble and insoluble fractions when cells were induced at 37 °C. PLZFBTB/POZ-transformed S9 cells that were induced at 28 °C resulted in the partitioning of the majority of the PLZF-BTB/POZ protein in the soluble protein fraction.

Purification—For protein purification, 300 ml of saturated PLZFBTB/POZ-transformed S9 culture was used to inoculate a total of 6 liters of Luria broth (50 ml of saturated culture/1 liter of Luria broth) that had been pre-warmed to 28 °C. Cells were grown to an A595 of 0.5 and induced with 0.5 mM isopropyl-1-thio-b-D-galactopyranoside, and growth was continued until the A595 reading plateaued. This typically took about 8 h. In our early purification attempts, we had noted that, once partially purified, the PLZF-BTB/POZ protein became insoluble when the pH was below 8.5. We subsequently used a purification strategy in which the protein was isolated in a buffer solution containing 40 mM Tris, pH 8.5. Following cell disruption by sonication in 100 ml of a buffer containing 40 mM Tris, pH 8.5, 100 mM NaCl, 1 mM bmercaptoethanol, and 100 mg/ml phenylmethylsulfonyl fluoride (LS buffer), the cell extract was centrifuged, and the supernatant was applied directly to a 10-ml Q-Sepharose column that had been preequilibrated with LS buffer. The column was washed with 10 column volumes of LS buffer containing 0.25 M NaCl, and protein was eluted with a linear gradient of salt (0.25– 0.4 M NaCl) in the same buffer. The fractions containing PLZF-BTB/POZ were pooled and adjusted to 0.75 M (NH4)2 SO4 prior to application to a 10-ml phenyl-Sepharose column, which had been pre-equilibrated with LS buffer containing 0.75 M (NH4)2 SO4. After washing the column with 10 column volumes of LS buffer containing 0.75 M (NH4)2 SO4, protein was eluted with a linear gradient of salt (0.75– 0 M (NH4)2 SO4) in LS buffer. Peak fractions containing PLZF-BTB/POZ protein were pooled and concentrated to about 2 ml by ultracentrifugation using a Centriprep-10 microconcentrator (Amicon Inc.). The concentrated protein was purified further using size exclusion chromatography on a Superdex-200 FPLC column (Pharmacia Biotech Inc., 16 3 600 mm). Peak fractions containing PLZF-BTB/POZ protein were concentrated to about 50 mg/ml by centrifugation using a Centricon-10 microconcentrator (Amicon Inc.) in LS buffer. 50-ml protein aliquots were finally frozen at 270 °C and thawed for use as needed. We typically obtained about 60 mg of purified protein from one 6-liter preparation of bacterial culture. Protein concentration was determined spectroscopically by absorbance at 280 nm using a molar extinction coefficient of 7450 M21 cm21 predicted from the amino acid sequence (29). The identity of the purified PLZF-BTB/POZ protein was confirmed by MALDI mass spectroscopy

27326

Characterization of the BTB/POZ Domain from PLZF

and N-terminal sequencing (Wistar Institute Protein Microchemistry Facility). Analytical Ultracentrifugation—The PLZF-BTB/POZ protein was dialyzed against a buffer containing 40 mM boric acid, pH 8.5, and 100 mM NaCl prior to analysis. Sedimentation equilibrium experiments were carried out at 4 °C in a Beckman XL-A analytical ultracentrifuge with two- or six-sector cells using three rotor speeds (20,000, 30,000, and 42,000 rpm) and at three different protein concentrations (0.2, 0.1, and 0.05 mg/ml). After samples had reached equilibrium (typically after 15 h), they were scanned at 280 and 230 nm. Data were analyzed with a nonlinear least squares fitting program based on IGRO-PRO (provided by P. Hensley, Smith Kline Beacham Pharmaceuticals) using a partial specific volume of 0.7347 ml/g calculated from the protein amino acid composition (30). Differential Scanning Calorimetry—DSC measurements were performed using a MicroCal MCS Calorimeter. Protein solutions were dialyzed against a buffer containing 40 mM boric acid, pH 8.5, and 100 mM NaCl prior to loading the sample cell. An aliquot of the dialysate was used in the reference cell. The protein concentration was 1.0 mg/ml (70 mM). All scans were from 10 to 90 °C at a scan rate of 90 degrees/h. ORIGIN software (MicroCal, Inc.) was used for data analysis, which involved subtracting a buffer base line from the raw data and then defining a base line using a progress curve fitted to the end of the transition. The constructed base-line was then subtracted, and the data were curve-fitted using standard models to determine Tm, calorimetric heat change, van’t Hoff heat change, and the heat capacity associated with the thermal unfolding. Dynamic Light Scattering—All measurements were carried out on a DynaPro-801 molecular sizing instrument at 20 °C. Samples were typically run at a concentration of 4 mg/ml (280 mM) and pre-filtered through a 0.02-mm membrane filter (Whatman, Anodisc 13) prior to analysis. Sample buffer typically contained 50 mM Tris-HCl, pH 8.5, 50 mM NaCl, 1 mM b-mercaptoethanol. Circular Dichroism Spectroscopy—CD spectra were carried out on a Jasco J-720 spectropolarimeter at 25 °C. The far-UV CD spectra were recorded using a 100-ml cell containing a 0.2-mm path length. The sample was typically at a concentration of 1 mg/ml (67 mM) in a buffer containing 20 mM boric acid, pH 8.5, 50 mM NaCl. Spectra were analyzed using the SOFTSPEC software supplied by the manufacturer. Equilibrium Denaturation—Denaturation experiments were performed by incubating individual samples of the PLZF-BTB/POZ protein in 20 mM boric acid, pH 8.5, 50 mM NaCl with increasing amounts of guanidine hydrochloride. These experiments were performed at two different protein concentrations, 1.87 mg/ml (125 mM) and 0.314 mg/ml (21 mM), utilizing sample cells with 0.2- and 1.0-mm path lengths, respectively. To measure the degree of unfolding, samples were analyzed at a wavelength of 222 nm using CD spectroscopy. The fraction of unfolded protein (fu) as a function of guanidine hydrochloride concentration was calculated with the equation f 5 ([u222]obs 2 [u222]f)/([u222]u 2 [u222]f), where [u222]obs is the molar helical ellipticity at a particular guanidine hydrochloride concentration and [u222]f and [u222]u are the molar helical ellipticity of the fully folded protein (in the absence of denaturant) and of the fully unfolded protein (high guanidine hydrochloride concentrations), respectively. [u222]f and [u222]u were obtained at the base lines of the transition curves, at which [u222]obs became relatively invariant at changing guanidine hydrochloride concentrations. The Gibbs free energy for each of the partially unfolded states was calculated assuming a two-state dimer denaturation as supported by the DSC results described above. Ku for unfolding was calculated using the equation Ku 5 2Pt[f 2u/(1 2 fu)], were Pt is the total protein concentration and fu is the fraction of unfolded protein at a particular guanidine hydrochloride concentration. The Gibbs free energy for unfolding at a particular guanidine hydrochloride concentration was then calculated with the equation DG 5 2RT ln(Ku), and the Gibbs free energy for protein unfolding in water (DGU(H2O)) was calculated by extrapolating to a value of DG at zero guanidine hydrochloride (31). Crystallization—Crystallization screens employed the vapor diffusion crystallization technique using 24-well culture plates (Linbro) (32). Typically, a 2-ml solution containing protein, salts, buffer, and precipitating agent were equilibrated against a 1-ml reservoir containing salts, buffer, and precipitating agent at twice the concentration contained in the protein drop. We employed three broad factorial crystallization screens (33–35). Each of the factorial screens were conducted at two temperatures, 20 and 4 °C. Crystallographic Data Collection and Processing—Diffraction data were collected on an R-AXIS II image plate with an MSC/YALE focusing mirror system using CuKa radiation from a Rigaku RU-200 x-ray

FIG. 2. Purification of the recombinant BTB/POZ domain of PLZF. A, an SDS-polyacrylamide (17.5%) gel is shown. Lane 1, molecular mass standards; lane 2, uninduced PLZF-BTB/POZ-transformed cells; lane 3, induced PLZF-BTB/POZ-transformed cells; lane 4, peak fraction of PLZF-BTB/POZ protein from chromatography on Q-Sepharose; lane 5, peak fraction of PLZF-BTB/POZ protein from chromatography on phenyl-Sepharose; and lane 6, concentrated PLZF-BTB/POZ protein (40 mg/ml) after size exclusion chromatography using a Superdex-200 FPLC column. B, size-exclusion chromatography using a Superdex-200 FPLC column in 40 mM Tris-HCL, pH 8.5, 100 mM NaCl, and 1 mM dithiothreitol. Elution volumes for protein standards are indicated with a dotted line. The PLZF-BTB/POZ protein elutes between the 17 and 44 kDa protein standards. generator operating at 100 mA and 50 kV. Data were processed with the programs DENZO and SCALEPACK (36). RESULTS

Purification of the BTB/POZ Domain from the PLZF Protein—We have successfully overexpressed and purified to homogeneity the BTB/POZ region from the PLZF protein (residues 1–118) in bacteria using a PQE30 expression plasmid in S9 cells (Fig. 2A). The protein preparation yields about 10 mg of purified protein per liter of bacterial cell culture. The availability of large amounts of purified recombinant protein has allowed us to study the biochemical and biophysical properties of this highly conserved BTB/POZ region from the PLZF protein. Oligomerization Properties of the BTB/POZ Domain from the PLZF Protein—Several biochemical/biophysical techniques, including gel filtration chromatography, dynamic light scattering, and equilibrium analytical ultracentrifugation, show that the recombinant PLZF-BTB/POZ protein exists as a stable dimeric species under near physiological conditions.

Characterization of the BTB/POZ Domain from PLZF

27327

FIG. 3. Equilibrium sedimentation of the recombinant BTB/ POZ domain of PLZF. The protein at an initial concentration of 15 mM in a buffer containing 20 mM boric acid, pH 8.5, 100 mM NaCl was analyzed by equilibrium analytical ultracentrifugation. Samples were centrifuged at 30,000 rpm at 4 °C for 20 h in a Beckman XL-I analytical ultracentrifuge. Data were analyzed in terms of a single sedimenting species as described previously (37) using a value of 0.7347 for the partial specific volume, computed by the method of Cohn and Edsall as described elsewhere (30). The fitted value for the molecular mass was 29,956 6 593 Da.

Size exclusion chromatography of the PLZF-BTB/POZ protein at a concentration above 2 mg/ml (140 mM) shows that it elutes between the horse myoglobin (17.0 kDa) and the chicken ovalbumin (44.0 kDa) protein standards (Fig. 2B). A plot of the logarithm of the molecular mass of the protein standards against the elution volume predicts that the molecular mass of the PLZF-BTB/POZ species is 30 kDa. This predicted molecular mass is most consistent with a dimeric state for the PLZF-BTB/ POZ protein since the molecular mass for a PLZF-BTB/POZ dimer calculated from the protein sequence is 29.97 kDa. Analysis of the PLZF-BTB/POZ protein by dynamic light scattering, using a DynaPro-801 molecular sizing instrument, shows that the protein has a Stokes radius of 2.7 nm, corresponding to a globular molecular mass of 32 kDa. As with the gel filtration analysis, this experiment suggests that the PLZFBTB/POZ protein exists as a dimer. To obtain an estimate of the molecular weight of the multimeric PLZF-BTB/POZ species that is independent of molecular shape and to investigate potential monomer-dimer equilibria, we performed equilibrium sedimentation experiments using analytical ultracentrifugation (Fig. 3). We performed these experiments in a buffer that was relatively transparent at 230 nm so that we could obtain information in the nanomolar concentration range of protein. Equilibrium sedimentation experiments were carried out at three rotor speeds (20,000, 30,000, and 42,000 rpm) and at three different initial protein concentrations (0.2, 0.1, and 0.05 mg/ml). The data from each run were fitted to a single molecular weight species yielding an average molecular mass of 29,956 6 593 Da over a protein concentration range of 33 mM to 200 nM. A representative curve fit at a rotor speed of 30,000 rpm and at a protein concentration of 0.05 mg/ml is shown in Fig. 3. This experiment shows that the PLZF-BTB/POZ domain is dimeric above a concentration of 200 nM. Taken together, gel filtration, dynamic light scattering, and sedimentation analytical ultracentrifugation experiments show that the predominant oligomeric state for the PLZF-BTB/ POZ protein is dimeric above a concentration of 200 nM under near physiological conditions.

FIG. 4. DSC scans of the BTB/POZ domain of PLZF. A, DSC scan raw data at a protein concentration of 1.0 mg/ml in buffer containing 40 mM boric acid, pH 8.5, 100 mM NaCl. —, initial sample heating; . . . ., second sample heating. B, base-line subtracted data (—) fitted to a single two-state transition (. . . .) using the ORIGIN software (MicroCal, Inc.). All data were collected using a scan rate of 90 °C/h.

Stability of the BTB/POZ Domain from the PLZF Protein—We performed both DSC and guanidine hydrochlorideinduced unfolding that was monitored by CD spectroscopy to investigate the stability of the PLZF-BTB/POZ dimer. DSC experiments showed that the PLZF-BTB/POZ fragment undergoes a single irreversible two-state transition with precipitation of the sample (Fig. 4A). The protein is surprisingly heat stable with an estimated Tm of 70.4 °C and an estimated enthalpy change of 48 kcal/mol (Fig. 4B). The ratio of calorimetric heat change over the van’t Hoff heat change was approximately 0.47, suggesting that the protein dimer undergoes a single coupled two-state transition during unfolding. The PLZF-BTB/POZ protein shows a strong mean residue ellipticity at 222 nm, suggesting a significant helical content. The ellipticity at 222 nm was used to evaluate unfolding of the PLZF-BTB/POZ protein in the presence of the denaturant guanidine hydrochloride. Varying concentrations of guanidine hy-

27328

Characterization of the BTB/POZ Domain from PLZF

FIG. 5. Guanidine hydrochloride-induced unfolding of the BTB/POZ domain of PLZF. Molar residue ellipticity at 222 nm ([u222]s) at various guanidine hydrochloride concentrations were measured in a buffer containing 20 mM boric acid, pH 8.5, 100 mM NaCl, 1 mM b-mercaptoethanol and converted to the fraction of unfolded protein (fu) at two different protein concentrations, 125 mM (f) and 21 mM (M). Data points were fitted to a sigmoidal curve.

drochloride were added, and the mean residue ellipticity was monitored at 222 nm. This analysis shows that unfolding is sigmoidal and single phase (Fig. 5). Moreover, the midpoints of the transition curves increased with increasing protein concentrations, further supporting the DSC experiments and suggesting a coupled two-state unfolding transition from folded dimer to unfolded monomers. The denaturation profiles of the PLZFBTB/POZ protein were used to calculate an average DGU(H2O) of 12.8 6 0.4 kcal/mol. Comparable Gibbs free energies for unfolding have been observed for other proteins of about the same size that form stable homodimers, including HIV-1 protease, and Trp aporepressor (31). Taken together the DSC and CD measurements show that the PLZF-BTB/POZ domain of PLZF forms an extremely stable, potentially intertwined dimer. Secondary Structure Content of the BTB/POZ Domain from the PLZF Protein—CD spectra of the PLZF-BTB/POZ domain shows that the protein has a high degree of secondary structure (Fig. 6). Analysis of the spectra using the SOFTSPEC software supplied by the manufacturer shows an excellent agreement with a protein containing 45% helix, 20% b-sheet, 11% turn, and 24% random coil. These experimental values are consistent with secondary structure predictions suggesting that about 50% of the BTB/POZ domain is helical (27). Crystallization and Diffraction Properties of PLZF-BTB/ POZ Crystals—The crystallization trials for the PLZF-BTB/ POZ protein produced several different crystal forms at 20 °C; however, none of these crystals diffracted x-rays to beyond 10 Å resolution. We obtained two crystal forms at 4 °C that were small but nicely shaped. We optimized crystal growth conditions for one of these forms and were able to reproducibly prepare crystals of typical size 0.2 3 0.2 3 0.5 mm in size. The crystals were prepared using 2-ml hanging drops containing 10 mg/ml PLZF-BTB/POZ protein, 8% isopropyl alcohol, 600 mM MgCl2, 50 mM Tris, pH 8.5, 50 mM HEPES, pH 7.5, equilibrated over a reservoir containing 2 times the concentration of salts, buffer, and precipitating agent. Crystals were transiently transferred (for about 5 min) to a harvest solution composed of salts, buffer, and precipitating agent at the same concentrations as the reservoir solution with the addition of 25% glycerol to facilitate x-ray data collection at cryogenic temperature (2170 °C). The diffraction quality of the crystals show that they are well ordered, showing strong diffraction to beyond 2.3 Å resolution, indicating that they would be suitable for structure determination using x-ray crystallographic techniques. A native diffraction data set (2.25 Å, Rsym 5 6.8%) was collected from a crystal at 2170 °C (Table I). Analysis and reduction of the x-ray data using the DENZO and SCALE-

FIG. 6. CD spectra of the BTB/POZ domain of PLZF. A, CD spectra of purified protein was carried out in a buffer containing 20 mM boric acid, pH 8.5, 50 mM NaCl at a protein concentration of 1 mg/ml (67 mM). B, a calculated CD spectra for a protein containing 45% helix, 20% b-sheet, 11% turn, and 24% random coil (- - - -) is plotted with the observed spectra (—) with the residual difference spectra (. . . .). TABLE I Data collection statistics for the native PLZF-BTB/POZ crystals Unit cell dimensions (Å) Space group Resolution limit (Å) Unique reflections Overall redundancy Overall I/Sigma1 Overall completeness Rmergea

a 5 38.8, b 5 77.7, c 5 85.3 I222 or I212121 2.25 6411 6.3 (6.2) 32.2 (17.2) 99.9% (100%) 0.068 (0.189)

a Rmerge 5 ShIh 2 ?Ih?/ShIh, where Ih is the intensity of reflection h, and Sh is the sum over all reflections. The value in parenthesis is for the last resolution shell (2.32–2.25 Å).

PACK programs (36) shows that the crystal forms in the space group I222 or I212121 with a 5 38.8, b 5 77.7, and c 5 85.3 Å and contain 1 protein subunit per asymmetric unit, with approximately 40% solvent content in the crystals. DISCUSSION

Oligomerization Properties of Proteins Containing a BTB/ POZ Domain—The biochemical/biophysical properties of the BTB/POZ region from the PLZF protein presented in this study show that it forms a highly stable homodimer. We have established an upper limit of 200 nM as a dissociation constant for the dimer, strongly suggesting that the PLZF protein exists as a dimer in vivo. Consistent with this is the apparent instability of the BTB/POZ monomer as determined by DSC and equilibrium denaturation experiments. The extensive homology within the BTB/POZ region of homologous proteins (Fig. 1) suggest that they also form dimers in vivo. This is supported by several in vivo and in vitro studies showing that the BTB/POZ domain from a variety of proteins mediate homodimerization and in some cases heterodimerization. Specifically, Laski and co-workers (14) have employed GST fusion proteins and cross-linking experiments to show that the BTB/POZ region of the Drosophila bric a brac protein mediates dimerization in vitro. These investigators also localize the dimerization surface to the N-terminal 51 residues of the BTB/ POZ domain and show that mutation of several highly conserved residues in this region disrupt the proteins oligomerization properties. Bardwell and Treisman (2) also have used coimmunoprecipitation experiments to demonstrate that the BTB/POZ domains of both the human ZID and Drosophila Ttk

Characterization of the BTB/POZ Domain from PLZF proteins can form homodimers in vitro. Moreover, they demonstrate that the BTB/POZ domain of the Drosophila Ttk protein form homodimers as well as heterodimers with the BTB/POZ region from the Drosophila GAGA protein. Taken together, a key property of the BTB/POZ domain appears to be to direct formation of particular homo and heterodimeric protein complexes. Involvement of the BTB/POZ Domain in PLZF-mediated APL—The PLZF-RARa fusion receptor generated as a result of a t(11;17) chromosomal translocation that occurs in a small subset of APL patients has been shown to display a dominantnegative effect against retinoic acid receptor function (18). Recently, Dong et al. (24) have shown that the BTB/POZ region of the PLZF protein mediates this activity. These authors also show that the PLZF-RARa fusion protein could heterodimerize in vitro with wild-type PLZF protein, suggesting that the BTB/ POZ domain may play a significant role in leukemogenesis by antagonizing not only the retinoid receptors but also PLZF and possible other BTB/POZ-domain containing regulators. Taken together, the dimerization properties of the PLZF BTB/POZ domain appears to be strongly linked to the molecular pathogenesis of APL. The studies presented here have contributed substantially to our understanding of the dimerization properties of the PLZF protein. The structure of the PLZF BTB/POZ region that will be afforded by the crystals obtained in this study will provide the high resolution chemical details underlying dimer stability and specificity. This information will provide a framework for understanding the structural basis underlying PLZF-RARa-mediated APL and for understanding the function and dimerization properties of other BTB/POZ-containing proteins. Acknowledgments—We thank A. Zelent for providing the PLZF cDNA (Leukemia Research Fund, Institute of Cancer Research); D. Christianson and laboratory (Department of Chemistry, the University of Pennsylvania) for use of the R-AXIS II image plate detector system for x-ray diffraction data collection; D. Speicher and S. Harper (Wistar Institute) for performing the DSC analysis and for helpful discussions; B. Degrado and J. Lear (Department of Biochemistry and Biophysics, the University of Pennsylvania Medical School) for performing the initial analytical ultracentrifugation studies; and L. Otvos (Wistar Institute) for access to the laboratory spectropolarimeter. This work was supported by a grant from the Leukemia Research Foundation to R. M. REFERENCES 1. Zollman, S., Godt, D., Prive, G. G., Couderc, J.-L., and Laski, F. (1994) Proc. Natl. Acad. Sci. U. S. A. 91, 10717–10721 2. Bardwell, V. J., and Treisman, R. (1994) Genes Dev. 8, 1664 –1677 3. Numoto, M., Niwa, O., Kaplan, J., Wong, K. K., Merrell, K., Kamiya, K., Yanagihara, K., and Calame, K. (1993) Nucleic Acids Res. 21, 3767–3775 4. Albagli, O., Dhordain, P., Deweindt, C., Lecocq, G., and Leprince, D. (1995) Cell Growth Differ. 6, 1193–1198

27329

5. DiBello, P. R., Withers, D. A., Bayer, C. A., Fristrom, J. W., and Guild, G. M. (1991) Genetics 129, 385–397 6. Dorn, R., Krauss, V., Reuter, G., and Saumweber, H. (1993) Proc. Natl. Acad. Sci. U. S. A. 90, 11376 –11380 7. Godt, D., Couderc, J.-L., Cramton, S. E., and Laski, F. A. (1993) Development 119, 799 – 812 8. Harrison, S. D., and Travers, A. A. (1990) EMBO J. 9, 207–216 9. Tsukiyama, T., Becker, P. B., and Wu, C. (1994) Nature 367, 525–532 10. Xiong, W. C., and Montell, C. (1993) Genes Dev. 7, 1085–1096 11. Xue, F., and Cooley, L. (1993) Cell 72, 681– 693 12. Lemons, R. S., Keller, S., Gietzen, D., Dufner, J., Rebentisch, M., Feusner, J., and Eilender, D. (1995) J. Pediatr. Hematol. Oncol. 17, 198 –210 13. Warrell, J., R. P., DeThe, H., Wang, Z.-Y., and Degos, L. (1993) N. Engl. J. Med. 329, 177–187 14. Chen, W., Zollman, S., Couderc, J. L., and Laski, F. A. (1995) Mol. Cell. Biol. 15, 3424 –3429 15. Rowley, J. D., Golomb, H. M., and Dougherty, C. (1977) Lancet 1, 549 –550 16. Zelent, A. (1994) Br. J. Haematol. 86, 451– 460 17. Chen, Z., Brand, N. J., Chen, A., Chen, S.-J., Tong, J.-H., Wang, Z. Y., Waxman, S., and Zelent, A. (1993) EMBO J. 12, 1161–1167 18. Chen, Z., Guidez, F., Rousselot, P., Agardir, A., Chen, S.-J., Wang, Z. Y., Degos, L., Zelent, A., Waxman, S., and Chomienne, C. (1994) Proc. Natl. Acad. Sci. U. S. A. 91, 1178 –1182 19. Licht, J. D., Chomienne, C., Goy, A., Chen, A., Scott, A. A., Head, D. R., Michaux, J. L., Wu, Y., DeBlasio, A., Jr., W. H. M., Zelenetz, A. D., Willman, C. L., Chen, Z., Chen, S.-J., Zelent, A., Macintyre, E., Veil, A., Cortes, J., Kantarjian, H., and Waxman, S. (1995) Blood 85, 1083–1094 20. Guidez, F., Huag, W., Tong, J.-H., Dubois, C., Balitrand, N., Waxman, S., Michaux, J. L., Martial, P., Degos, L., Chen, Z., and Chomienne, C. (1994) Leukemia (Baltimore) 8, 312–317 21. Pengue, G., Calabro, V., Baroli, P. C., Pagliuca, A., and Lania, L. (1994) Nucleic Acids Res. 22, 2908 –2914 22. Freemont, P. S. (1993) Zinc Finger Proteins in Oncogenesis (Sluyser, M. A. B., Brinkmann, A. O., and Blankenstein, R. A., eds) The New York Acadamy of Science, New York 23. Licht, J. D., Shaknovich, R., English, M. A., Melnick, A., Li, J.-Y., Reddy, J. C., Dong, S., Chen, S.-J., Zelent, A., and Waxman, S. (1996) Oncogene 12, 323–336 24. Dong, S., Zhu, J., Reid, A., Strutt, P., Guidez, F., Zhong, H.-J., Wang, Z.-Y., Lict, J., Waxman, S., Chomienne, C., Chen, Z., and Zelent, A. (1996) Proc. Natl. Acad. Sci. U. S. A. 93, 3624 –3629 25. Bavisotto, L., Kaushansky, K., Lin, N., and Hromas, R. (1991) J. Exp. Med. 174, 1097–1101 26. Nguyen, H. Q., Hoffman-Liebermann, B., and Liebermann, D. A. (1993) Cell 72, 197–209 27. Reid, A., Gould, A., Brand, N., Cook, M., Strutt, P., Li, J., Licht, J., Waxman, S., Krumlauf, R., and Zelent, A. (1995) Blood 86, 4544 – 4552 28. Bujard, H., Gentz, R., Lanzer, M., Stuber, D., Muller, M., Ibrahmi, I., Hauptle, M. T., and Dobberstein, B. (1987) Methods Enzymol. 155, 416 – 433 29. Pace, C. N., Vajdos, F., Fee, L., Grimsley, G., and Gray, T. (1995) Protein Sci. 4, 2411–2423 30. Laue, T. M., Shah, B. D., Ridgeway, T. M., and Pelletier, S. L. (1992) in Analytical Ultracentrifugation in Biochemistry and Polymer Science (Harding, S. E., Rowe, A. J., and Horton, J. C., eds) pp. 90 –125, Royal Chemical Society, Cambridge, UK 31. Neet, K. E., and Timm, D. E. (1994) Protein Sci. 3, 2167–2174 32. McPherson, A. (1982) The Preparation and Analysis of Protein Crystals, John Wiley & Sons, Inc., New York 33. Cudney, R., Patel, S., Weisgraber, K., Newhouse, Y., and McPherson, A. (1994) Acta Crystallogr. Sec. D 50, 414 – 423 34. Jancarik, J., and Kim, S.-H. (1991) J. Appl. Crystalogr. 24, 409 – 411 35. Scott, W. G., Finch, J. T., Grenfell, R., Fogg, J., Smith, J., Gait, M. J., and Klug, A. (1995) J. Mol. Biol. 250, 327–332 36. Gewirth, D., Otwinowski, Z., and Minor, W. (1993) The HKL Version 1.0 Manual, 3rd Ed., Yale University Press, New Haven, CT 37. Brooks, I., Watts, D. G., Soneson, K. K., and Hensley, P. (1994) Methods Enzymol. 240, 459 – 478