research communications
Saccharomyces cerevisiae
of cytoplasmic acetoacetyl-CoA thiolase fromaHefei National Laboratory for Physical Sciences at the Microscale and School of Life Sciences, University of Science and Technology of China, Hefei, Anhui 230026, People's Republic of China
*Correspondence e-mail: zlzhu63@ustc.edu.cn, lwniu@ustc.edu.cn
Thiolases are vital enzymes which participate in both degradative and biosynthetic pathways. Biosynthetic thiolases catalyze carbon–carbon bond formation by a Claisen condensation reaction. The cytoplasmic acetoacetyl-CoA thiolase from Saccharomyces cerevisiae, ERG10, catalyses carbon–carbon bond formation in the mevalonate pathway. The structure of a S. cerevisiae biosynthetic thiolase has not previously been reported. Here, crystal structures of apo ERG10 and its Cys91Ala variant were solved at resolutions of 2.2 and 1.95 Å, respectively. The structure determined shows that ERG10 shares the characteristic thiolase superfamily fold, with a similar active-site architecture to those of type II thiolases and a similar binding pocket, apart from Ala159 at the entrance to the pantetheine-binding cavity, which appears to be a determinant of the poor binding ability of the substrate. Moreover, comparative binding-pocket analysis of molecule B in the of the apo structure with that of the CoA-bound complex of human mitochondrial acetoacetyl-CoA thiolase indicates the canonical binding mode of CoA. Furthermore, the revealed in a structural comparison of molecule A with the CoA-bound form raise the possibility of conformational changes that are associated with substrate binding.
Keywords: acetoacetyl-CoA thiolase; Claisen condensation; mevalonate pathway; X-ray crystallography; Saccharomyces cerevisiae.
PDB references: acetyl-CoA acetyltransferase, native, 5xz5; mutant, 5xyj
1. Introduction
Thiolases are vital ubiquitous enzymes which catalyze the formation of carbon–carbon bonds via a Claisen condensation reaction. Thiolases catalyze the acetylation of a reactive cysteine residue followed by a condensation reaction, which is the major enzymatic route of fatty-acid, polyketide and steroid synthesis (Haapalainen et al., 2006). Thiolases have been divided into degradative thiolases (EC 2.3.1.16) and synthetic thiolases (EC 2.3.1.9) based on their biological functions and substrate specificity (Modis & Wierenga, 1999). Degradative thiolases, which are also termed 3-ketoacyl-CoA thiolases, exhibit broad chain substrate specificity, while biosynthetic thiolases or acetoacetyl-CoA thiolases are highly specific for short-chain substrates such as acetyl-CoA and acetoacetyl-CoA (Kursula et al., 2005).
The Saccharomyces cerevisiae degradative peroxisomal thiolase is involved in fatty-acid β-oxidation and the implications of its structure for substrate binding and have been reported (Mathieu et al., 1997). In contrast, the cytoplasmic acetoacetyl-CoA thiolase from S. cerevisiae (ERG10) catalyses the first step of the mevalonate pathway by a reversible nondecarboxylative condensation reaction in which acetoacetyl-CoA is synthesized from two molecules of acetyl-CoA (Jiang et al., 2008). All isoprene-containing compounds such as dolichol and isopentenylated adenosine in tRNAs are products of this pathway. Ergosterol was found to be the major sterol product of this pathway (Hiser et al., 1994).
Various structures of human and bacterial thiolases in complexes with their substrates have been resolved to investigate the binding mode of the substrates and their catalytic mechanisms, but the structure of a synthetic thiolase from a lower eukaryote such as S. cerevisiae as yet remains unexplored. After we had resolved the of the wild-type thiolase, we attempted to obtain the complex of the Cys91Ala mutant thiolase with CoA by co-crystallization, but no obvious density for CoA was observed. The current work concludes that ERG10 shares a similar overall structure with biosynthetic thiolases and also possesses a conserved Cys–Asn–His catalytic triad in addition to another Cys residue that acts as a base. Comparative analysis of the binding pockets between molecule B of the and human mitochondrial acetoacetyl-CoA thiolase (T2) reveals a highly typical interaction behaviour for ERG10 and CoA, with the exception of Ala159. The clashes that result on docking CoA into molecule A of the reveal conformational changes on CoA binding.
2. Materials and methods
2.1. Recombinant protein production
The gene encoding native ERG10 was amplified by PCR from S. cerevisiae genomic DNA and cloned into pET-22bN vector using NdeI and XhoI sites to encode recombinant protein with a hexahistidine tag at the N-terminus. The Cys91Ala mutant was obtained by site-directed mutagenesis utilizing the resulting plasmid as a template, as reported by Sato et al. (1998). Both recombinant plasmids were transformed into Escherichia coli BL21 (DE3) cells for expression. Protein expression was induced at 289 K for approximately 20 h by adding 0.5 mM isopropyl β-D-1-thiogalactopyranoside (IPTG). Bacterial cells were harvested by centrifugation at 6080g for 6 min, resuspended in buffer A (50 mM Tris–HCl pH 8.0, 500 mM NaCl, 5% glycerol) and disrupted by sonication via an ultrasonicator (Qsonica, USA). Cell debris was removed by ultracentrifugation (20 000g for 30 min at 277 K) and the recombinant protein was purified from the supernatant using a two-step purification protocol consisting of Ni2+-affinity (Ni2+–NTA column; GE Healthcare, USA) and (HiLoad 16/60 Superdex 200 column; GE Healthcare, USA). The same procedure was used to purify the Cys91Ala mutant. Prior to crystallization, both protein samples were diluted to 10 mg ml−1 in buffer B (20 mM Tris–HCl pH 8.0, 200 mM NaCl, 5% glycerol). The mutant protein was incubated at room temperature with CoA at a molar ratio of 1:10 for 1 h. Macromolecule-production information is summarized in Table 1.
‡The XhoI site is underlined. §The His6 tag is in shown in italics. The amino acid in bold indicates the position of the mutation. |
2.2. Crystallization
The hanging-drop vapour-diffusion method was used for initial crystallization screening utilizing various conditions. Both the native and the mutant protein crystallized in several conditions. The drops were prepared using 1.0 µl protein sample and an equal volume of crystallization reservoir solution and were equilibrated against 100 µl precipitant solution as described in Table 2.
|
2.3. X-ray data collection and processing
The crystal was fished out with a clean loop and flash-cooled in liquid nitrogen at 100 K using reservoir solution containing 20%(v/v) glycerol as a cryoprotectant. X-ray diffraction data sets for the wild-type and mutant thiolases were collected on beamlines BL17U1 and BL18U1 at Shanghai Synchrotron Radiation Facility (SSRF). The available data sets were indexed, integrated and scaled using the HKL-2000 program suite (Otwinowski & Minor, 1997). Data-collection and processing statistics are detailed in Table 3.
|
2.4. Structure solution and refinement
MOLREP from the CCP4 program suite (Vagin & Teplyakov, 2010) was used to solve the structure of ERG10 by the molecular-replacement method using a single monomer of the tetrameric human mitochondrial thiolase (PDB entry 2ib7; Haapalainen et al., 2007) as a search model. The final structure model was completed by alternating manual building in Coot (Emsley et al., 2010) and in REFMAC5 (Murshudov et al., 2011). An overall assessment of the model quality was made using the structure-validation program PROCHECK (Laskowski et al., 1993). The same procedure was also applied to resolve the structure of the Cys91Ala mutant. The and stereochemical statistics for the two structures are summarized in Table 4. The atomic coordinates and structure factors were deposited in the PDB as entries 5xz5 and 5xyj for the wild-type and mutant thiolases, respectively. Figures showing structures were prepared using PyMOL (DeLano, 2002).
‡Rfree is the same as Rwork but for a 5% subset of all reflections that were not used in crystallographic §Root-mean-square deviation from ideal values. ¶The categories were defined by PROCHECK. |
3. Results and discussion
3.1. Oligomerization
The apparent molecular weight of wild-type ERG10 in solution was determined by comparison with protein standards of precisely known molecular weights using a HiLoad 16/60 Superdex 200 column. The elution volumes of protein standards were used to calculate a standard curve. The elution volume (63.07 ml) of the predominant peak corresponds to an apparent molecular weight of 164 kDa for ERG10, which coincides with the result reported by Kornblatt & Rudney (1971) and demonstrates the tetrameric state of ERG10 in solution (Fig. 1).
3.2. Overall structure and active site
The Homo sapiens (Haapalainen et al., 2007) and cytoplastic thiolase (bCT) from Zoogloea ramigera (Modis & Wierenga, 2000) could be constructed by applying operations (Fig. 2a) and is shown not to be a crystallographic artefact by the tetrameric state in solution. In essence, the structure of ERG10 is similar to those of T2 and bCT, even to the level of details of the quaternary-structure organization. Each subunit of ERG10 consists of N-terminal, C-terminal and loop domains. The active site is constructed on the N-terminal and C-terminal domains and its identical architecture to those of T2 and bCT reveals that the catalytic mechanism of ERG10 is similar to those of type II thiolases. The loop domain plays multiple roles in shaping the binding pocket for the CoA moiety of the substrate, stabilizing the tetramer and covering the active-site cavity using a tetramerization loop and a covering loop, respectively (Fig. 2b). Although the overall structure of ERG10 is highly similar to those of T2 and bCT, it differs in the following three key aspects: (i) the presence of Ala159 at the entrance to the pantetheine-binding cavity, (ii) the presence of a disordered region in molecule A of the and (iii) the steric clashes that are observed on docking CoA into molecule A.
of the crystal contains two protomers, but a tetramer similar to those found in type II biosynthetic thiolases such as mitochondrial thiolase (T2) from3.3. Comparison of the binding pocket between ERG10 and T2
A high degree of sequence identity was found among ERG10, T2 and bCT. Sequence alignment of ERG10 with T2 and bCT helps us to determine the binding and catalytic residues in ERG10 (Fig. 3). Generally, the protein residues and their spatial orientations for binding the adenosine 3′-phosphate and pantetheine moieties of the CoA molecule are similar when ERG10 is superposed on T2 (Fig. 4a), suggesting that ERG10 has a similar CoA-binding mode. The β-mercaptoethylamine and pantothenic acid moieties of the CoA molecule were stabilized by Ser284 and His192 in T2. A crucial water molecule that hydrogen-bonds to His192 interacts with the pantetheine moiety via a hydrogen bond, as previously described by Haapalainen et al. (2007). In chain B of ERG10, Ser252 was found at the position corresponding to Ser284 in T2. A significant substitution of histidine by alanine (Ala159) was found in ERG10 at the position corresponding to His192 in T2 by sequence alignment. Consequently, the key water molecule is not observed at a similar position, which would further result in a loss of the hydrogen-bond interactions between the enzyme and substrate (Fig. 4b).
Apart from covering the active-site cavity, the covering loop also plays an important role in binding the pantetheine moiety of the CoA molecule by contributing another two residues to the binding pockets of T2 and bCT, which is in agreement with the residues (Leu151 and Met160) present at the same positions in the covering loop of ERG10. The surface-charge distributions of the covering loop are remarkably similar in T2 and bCT, but are significantly different in ERG10, with negative potential contributed by Ala159 and Glu156, respectively (Fig. 5).
Enzyme assays of T2 and bCT gave Km values of 8 µM (Haapalainen et al., 2007) and 15 µM (Masamune et al., 1989) for acetoacetyl-CoA, respectively. The higher Km value of 350 µM for ERG10 as determined by Kornblatt & Rudney (1971) indicates a poor ability to bind the acetoacetyl-CoA molecule, probably owing to the substitution of histidine by alanine (Ala159) in the binding pocket, which is in line with the findings of Kim & Kim (2014), who reported that a His156Ala mutation in RePhaA (a β-ketothiolase from Ralstonia eutropha) leads to a lower enzymatic activity compared with the wild type. The impact of the negatively charged of the covering loop on the interaction between ERG10 and substrates remains unclear.
3.4. Implications of the disordered region and steric clashes
A high B factor indicates a lower possibility of finding an atom at its mean position in a (Willis & Pryor, 1975). Residues in regions of high B factor are considered to be the result of dynamic disorder of the polypeptide chain or short-range or long-range disorder within the crystal (Deller et al., 2016). During tracing and structure of the ERG10 polypeptide chains, we found that residues Ala235–Asn250 of subunit A of ERG10 are missing in the electron density (Fig. 6) and have higher B factors than the corresponding regions of subunit B of ERG10 and the two subunits of the Cys91Ala mutant (Fig. 7). Hence, this polypeptide fragment is inherently disordered and was removed from the final structural model of ERG10. Missing regions of electron density in several proteins are likely to perform important functions, as reported by Huber & Bennett (1983). This peptide fragment may provide Phe239 to bind the pantetheine moiety of the CoA molecule by hydrophobic interactions (Fig. 4b).
As shown in Fig. 8(a), structural superposition of the two subunits in the of ERG10 shows an ordered conformation in the corresponding region in subunit B, whereas a flexible conformation and obvious polypeptide-chain movements are observed at both ends of the disordered region in subunit A. Furthermore, superposition of subunit A of apo ERG10 and the structure of T2 in complex with a CoA molecule revealed significant structural differences between the binding pocket at the Cα atoms of His228 (2.0 Å from the Cα atom of Asp260 in T2) and Ser252 (5.5 Å from the Cα atom of Ser284 in T2). The differences in the positions of these two residues would lead to steric clashes with the adenine and pantetheine moieties of the CoA molecule, respectively (Fig. 8b), which raises the possible hypothesis that binding of the CoA molecule induces conformational changes, as has been commonly observed but which remains unreported in the thiolase superfamily.
Protein folding is affected by many factors, including pH, primary structure, ligand binding and crystal packing (Deller et al., 2016). The peptide fragment Ala235–Asn250 is involved in crystal packing in both the native protein and the Cys91Ala mutant (Fig. 9). The stable conformation of the peptide fragment (Fig. 8c) prompts us to speculate that the slightly tighter crystal packing in the mutant is perhaps a result of the CoA molecule binding but with differential occupancy. Further research is required to explore the impact of Ala159 on the affinity, and the structure of the ligand-bound mutant would be useful to test our hypothesis on the conformational changes involved in ligand binding.
Acknowledgements
We thank the staff at beamline BL17U1 of the Shanghai Synchrotron Radiation Facility (SSRF) and beamline BL18U1 of the National Facility for Protein Science (NFPS) at SSRF for assistance during data collection.
Funding information
This work was supported by the Chinese National Natural Science Foundation (grant Nos. U1632124, 31270770 and 31621002) and the Chinese Ministry of Science and Technology (grant Nos. 2017YFA0503600 and 2017YFA0504903).
References
DeLano, W. L. (2002). CCP4 Newsl. Protein Crystallogr. 40, 82–92. https://www.ccp4.ac.uk/newsletters/newsletter40/11_pymol.html. Google Scholar
Deller, M. C., Kong, L. & Rupp, B. (2016). Acta Cryst. F72, 72–95. CrossRef IUCr Journals Google Scholar
Emsley, P., Lohkamp, B., Scott, W. G. & Cowtan, K. (2010). Acta Cryst. D66, 486–501. Web of Science CrossRef CAS IUCr Journals Google Scholar
Haapalainen, A. M., Meriläinen, G., Pirilä, P. L., Kondo, N., Fukao, T. & Wierenga, R. K. (2007). Biochemistry, 46, 4305–4321. Web of Science CrossRef PubMed CAS Google Scholar
Haapalainen, A. M., Meriläinen, G. & Wierenga, R. K. (2006). Trends Biochem. Sci. 31, 64–71. Web of Science CrossRef PubMed CAS Google Scholar
Hiser, L., Basson, M. E. & Rine, J. (1994). J. Biol. Chem. 269, 31383–31389. CAS PubMed Google Scholar
Huber, R. & Bennett, W. S. (1983). Biopolymers, 22, 261–279. CrossRef CAS PubMed Google Scholar
Jiang, C., Kim, S. Y. & Suh, D.-Y. (2008). Mol. Phylogenet. Evol. 49, 691–701. CrossRef PubMed CAS Google Scholar
Kim, E.-J. & Kim, K.-J. (2014). Biochem. Biophys. Res. Commun. 452, 124–129. Web of Science CrossRef CAS PubMed Google Scholar
Kornblatt, J. A. & Rudney, H. (1971). J. Biol. Chem. 246, 4417–4423. CAS PubMed Google Scholar
Kursula, P., Sikkilä, H., Fukao, T., Kondo, N. & Wierenga, R. K. (2005). J. Mol. Biol. 347, 189–201. Web of Science CrossRef PubMed CAS Google Scholar
Laskowski, R. A., MacArthur, M. W., Moss, D. S. & Thornton, J. M. (1993). J. Appl. Cryst. 26, 283–291. CrossRef CAS Web of Science IUCr Journals Google Scholar
Masamune, S., Walsh, C., Sinskey, A. & Peoples, O. (1989). Pure Appl. Chem. 61, 303–312. CrossRef CAS Google Scholar
Mathieu, M., Modis, Y., Zeelen, J. P., Engel, C. K., Abagyan, R. A., Ahlberg, A., Rasmussen, B., Lamzin, V. S., Kunau, W. H. & Wierenga, R. K. (1997). J. Mol. Biol. 273, 714–728. CrossRef CAS PubMed Web of Science Google Scholar
Modis, Y. & Wierenga, R. K. (1999). Structure, 7, 1279–1290. Web of Science CrossRef PubMed CAS Google Scholar
Modis, Y. & Wierenga, R. K. (2000). J. Mol. Biol. 297, 1171–1182. Web of Science CrossRef PubMed CAS Google Scholar
Murshudov, G. N., Skubák, P., Lebedev, A. A., Pannu, N. S., Steiner, R. A., Nicholls, R. A., Winn, M. D., Long, F. & Vagin, A. A. (2011). Acta Cryst. D67, 355–367. Web of Science CrossRef CAS IUCr Journals Google Scholar
Otwinowski, Z. & Minor, W. (1997). Methods Enzymol. 276, 307–326. CrossRef CAS PubMed Web of Science Google Scholar
Robert, X. & Gouet, P. (2014). Nucleic Acids Res. 42, W320–W324. Web of Science CrossRef CAS PubMed Google Scholar
Sato, T., Kanai, Y. & Hoshino, T. (1998). Biosci. Biotechnol. Biochem. 62, 407–411. CrossRef CAS PubMed Google Scholar
Vagin, A. & Teplyakov, A. (2010). Acta Cryst. D66, 22–25. Web of Science CrossRef CAS IUCr Journals Google Scholar
Willis, B. T. M. & Pryor, A. W. (1975). Thermal Vibrations in Crystallography. Cambridge University Press. Google Scholar
© International Union of Crystallography. Prior permission is not required to reproduce short quotations, tables and figures from this article, provided the original authors and source are cited. For more information, click here.