Next Article in Journal
Towards Computational Modeling of Ligand Binding to the ILPR G-Quadruplex
Next Article in Special Issue
Thymosin α1 and Its Role in Viral Infectious Diseases: The Mechanism and Clinical Application
Previous Article in Journal
Simultaneous Quantification of Trehalose and Trehalose 6-Phosphate by Hydrophilic Interaction Chromatography/Electrospray Accurate Mass Spectrometry with Application in Non-Targeted Metabolomics
Previous Article in Special Issue
Evolutionary Couplings and Molecular Dynamic Simulations Highlight Details of GPCRs Heterodimers’ Interfaces
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Structural Analysis of Spermidine Synthase from Kluyveromyces lactis

1
Department of Biology Education, Kyungpook National University, 80 Daehak-ro, Buk-gu, Daegu 41566, Republic of Korea
2
Department of Biomedical Convergence Science and Technology, Kyungpook National University, 80 Daehak-ro, Buk-gu, Daegu 41566, Republic of Korea
3
Science Education Research Institute, Kyungpook National University, 80 Daehak-ro, Buk-gu, Daegu 41566, Republic of Korea
*
Author to whom correspondence should be addressed.
Molecules 2023, 28(8), 3446; https://doi.org/10.3390/molecules28083446
Submission received: 16 March 2023 / Revised: 7 April 2023 / Accepted: 11 April 2023 / Published: 13 April 2023
(This article belongs to the Special Issue Protein Structure, Function and Interaction)

Abstract

:
Spermidine is a polyamine molecule that performs various cellular functions, such as DNA and RNA stabilization, autophagy modulation, and eIF5A formation, and is generated from putrescine by aminopropyltransferase spermidine synthase (SpdS). During synthesis, the aminopropyl moiety is donated from decarboxylated S-adenosylmethionine to form putrescine, with 5′-deoxy-5′-methylthioadenosine being produced as a byproduct. Although the molecular mechanism of SpdS function has been well-established, its structure-based evolutionary relationships remain to be fully understood. Moreover, only a few structural studies have been conducted on SpdS from fungal species. Here, we determined the crystal structure of an apo-form of SpdS from Kluyveromyces lactis (KlSpdS) at 1.9 Å resolution. Structural comparison with its homologs revealed a conformational change in the α6 helix linked to the gate-keeping loop, with approximately 40° outward rotation. This change caused the catalytic residue Asp170 to move outward, possibly due to the absence of a ligand in the active site. These findings improve our understanding of the structural diversity of SpdS and provide a missing link that expands our knowledge of the structural features of SpdS in fungal species.

1. Introduction

Polyamines are ubiquitous components in most cells, among which diamine putrescine, triamine spermidine, and tetraamine spermine are widely found in living organisms [1,2,3,4]. Other polyamines such as thermospermine are found in thermophiles that survive at extremely high temperatures [5,6]. Polyamines bind easily to cellular polyanions; in Escherichia coli, almost half of putrescine and 90% of spermidine is complexed with cellular RNA [7,8]. Spermidine is a type of polyamine produced from a shorter chain putrescine, which is involved in various biological processes, including the regulation of membrane potential, inhibition of nitric oxide synthase (NOS), and the induction of autophagy [9]. In particular, spermidine suppresses protein acetylation by inhibiting the activity of acetyltransferases such as E1A-associated protein p300 (EP300) [10], which can rapidly induce autophagy by altering autophagic flux [10]. Spermidine also influences translation through eIF5A, which forms an uncommon amino acid hypusine on eIF5A by conjugating a lysine residue and the aminobutyl moiety from spermidine [11,12].
Aminopropyltransferases are involved in synthesizing polyamines in dependence of decarboxylated S-adenosylmethionine (dcSAM), which is converted from S-adenosylmethionine (SAM) by S-adenosylmethionine decarboxylase (SAMDC) [13,14,15,16]. In turn, dcSAM facilitates the catalytic reaction by transferring its aminopropyl moiety to a shorter-chain polyamine, resulting in a longer-chain polyamine, with methylthioadenosine (MTA) formed as a byproduct [17,18].
Spermidine synthase (SpdS; EC 2.5.1.16) is a major type of aminopropyltransferase that converts putrescine into spermidine [13,14,15]. The general reaction mechanism of SpdS has been well established: putrescine initiates nucleophilic attack on dcSAM, which donates an aminopropyl moiety to active site residues, such as Asp, Tyr, and Ser, to produce spermidine [1]. There are two types of enzymatic mechanisms for SpdS: ping-pong and sequential. SpdS in Glycine max (soybean) and E. coli follow a ping-pong mechanism [19,20], whereas SpdS in Thermotoga maritima, Plasmodium falciparum, Rattus rattus (rat), and Homo sapiens utilize a sequential mechanism [1,21,22,23].
SpdS consists of an N-terminal domain, which contains four β-strands, and a C-terminal domain containing a Rossmann-like fold [22]. In addition, SpdS contains a structural feature known as the gate-keeping loop, which is located in the vicinity of the entrance to the active site [22,24]. This loop is involved in recognizing the putrescine substrate, and mutational studies have revealed that several residues in this loop contribute toward substrate-binding and stabilization of the active site [25]. Since the first crystal structure of an aminopropyltransferase was reported in 2002 from the thermophilic anaerobic bacteria T. maritima [22], several other SpdS structures have been made available for several species including H. sapiens [1], Arabidopsis thaliana [26], P. falciparum [27], Helicobacter pylori [28], and E. coli [3]; however, our understanding of the structural features and evolutionary relationships of SpdS from fungal species remains limited.
In this study, we determined the crystal structure of the fungal Kluyveromyces lactis SpdS (KlSpdS) and compared its gate-keeping loop and active site with homologous structures. Structural analysis revealed several distinct conformational features in KlSpdS.

2. Results

2.1. Overall Structure

KlSpdS exists as a dimer in the asymmetric unit, and each monomer in the dimeric KlSpdS is positioned with two-fold symmetry (Figure 1A). The N- and C-terminal regions in each monomer are mainly involved in dimerization via interactions with β3 and α8-α9 of their partner molecules. Each KlSpdS monomer consists of three domains: an N-terminal domain (residues 4–66), a central catalytic core domain (residues 67–250), and a C-terminal domain (residues 251–292; Figure 1B, Supplemental Figure S1). The N-terminal domain includes six β-strands and is smaller than the catalytic core domain. The first two β-sheets of the N-terminal domain form a β-hairpin structure, followed by a four-stranded β-strands. Meanwhile, the catalytic core domain contains seven β-strands that form a Rossmann-like fold from β7 to β13 and seven α-helices. This canonical topology appears widely in nucleotide-binding enzymes and in class I MTases, which use dcSAM as a methyl moiety donor [24]. The C-terminal domain includes three α-helices (α8–α9) that mainly contribute toward dimer formation. The electrostatic surface representation of KlSpdS revealed a large cavity in the catalytic core domain between the N- and C-terminal domains (Figure 1C); the cavity was highly negatively charged, suggesting that positively charged dcSAM bind to putrescine in this active site pocket.

2.2. Gate-Keeping Loop

To elucidate the conformational diversity of the gate-keeping loop, the KlSpdS structure was superimposed with several homologous SpdS structures (Figure 2, Table 1). While the gate-keeping loop of H. sapiens SpdS (HsSpdS) sterically hindered the entrance of the active site, that of KlSpdS had an open conformation (Figure 2A). These conformational changes could be attributed to a short α-helix (α6) in the loop that forms on one side of the putrescine-binding region and may facilitate the accurate localization of putrescine in the active site. In KlSpdS, the α6 helix kinked outward by approximately 41.0° compared to HsSpdS and by approximately 34.6° compared to A. thaliana SpdS (AtSpdS; Figure 2B). The gate-keeping loop in Thermus thermophilus SpdS (TtSpdS) had a slightly different conformation compared to that of HsSpdS and AtSpdS (Figure 2C) and the α6 helix kinked inward by 31.1° compared to KlSpdS. Notably, the α6 helix of Thermotoga maritima SpdS (TmSpdS) was shorter than that of KlSpdS, suggesting that its longer gate-keeping loop could be more flexible (Figure 2D).
To assess conformational changes in the gate-keeping loop upon ligand binding, the KlSpdS structure was superimposed with the available structures of dcSAM complexed with HsSpdS, AtSpdS, P. falciparum SpdS (PfSpdS), and TcSpdS (Supplemental Figure S2, Table 1). The gate-keeping loop in the HsSpdS-dcSAM complex was not visible due to disordered (Supplemental Figure S2A), indicating that dcSAM could open the active site in HsSpdS by altering the conformation of the gate-keeping loop. Meanwhile, the α6 helix conformation of the AtSpdS-dcSAM complex differed by approximately 35.2° compared to KlSpdS (Supplemental Figure S2B). The conformations of apo-AtSpdS and the AtSpdS-dcSAM complex were highly similar, indicating that dcSAM binding has no significant effect on AtSpdS conformation. The α6 helix conformations of PfSpdS and TcSpdS complexed with dcSAM also differed from those of KlSpdS by approximately 35° (Supplemental Figure S2C,D).
Next, we investigated whether the gate-keeping loop conformation changed upon ligand binding in various species. No significant changes were observed in the gate-keeping loop conformation with binding of ligands such as MTA, adoDATO, 4MCA, putrescine, dcSAM, and spermidine in A. thaliana, T. thermophilus, T. maritima, or P. falciparum; however, changes were observed for H. sapiens (Supplemental Figure S3). TtSpdS-MTA, TmSpdS-adoDATA, and AtSpdS-4MCHA complexes shared similar gate-keeping loop conformations. Although the α6 helix induced fit upon ligand binding in TcSpdS, no conformational changes in the gate-keeping loop were observed in the other four species (Supplemental Figures S2 and S3). When we compared the structures of HsSpdS and PfSpdS in complex with putrescine, dcSAM, spermidine, and MTA, the gate-keeping loops exhibited almost the same conformation, except for those in the MTA complexes, which had a transition angle of 9.6° (Supplemental Figure S4). This might be attributed to the residues joining this region, which were nearly the same except for Ile201 in PfSpdS instead of Met178 in HsSpdS. However, since both amino acids have a non-polar character, the conformational differences might not be substantial.

2.3. Active Site

The catalytic residues Asp98, Asp167, and Asp 170 of KlSpdS were highly conserved in other SpdS structures (Figure 3). Asp98 captures the aminopropyl moiety of dcSAM and remains ready for the initiation of nucleophilic attack by putrescine. Asp167 plays a crucial role in the deprotonation of putrescine, while Asp170 is required for accurate putrescine binding. To examine the possible active site of KlSpdS, its structure was superposed with those of HsSpdS, AtSpdS, PfSpdS, and TcSpdS complexed with specific ligands (Figure 3A–D). Although Asp98 and Asp167 had conformations similar to the other structures, conformation of Asp170 was distinct, possibly due to changes caused by ligand binding.
Overall, most ligand-interacting residues were similar in the structures of HsSpdS and KlSpdS complexed with putrescine substrate, with the Tyr73, Asp167, and Ile240 residues in KlSpdS aligning especially well with the corresponding Tyr79, Asp173, and Ile246 residues in HsSpdS (Figure 4A). Putrescine generated four hydrogen bonds with the amino acids present in the active site, including three residues in the gate-keeping loop. With spermidine, most residues were well matched except for Ser174, Ser175, Asp176, and Try241 (Figure 4B), and it was stabilized by six of the seven possible hydrogen bonds between the gate-keeping loop and the active site. With the cofactor dcSAM, most residues aligned well except for Glu124, Pro180, and Leu184 (Figure 4C). Similarly, most residues were matched when bound to the MTA byproduct, except for Glu124, Pro180, and Leu184 (Figure 4D). In particular, Glu124 in KlSpdS corresponded to Asp118 in HsSpdS, both are negatively charged, suggesting that there would be no critical change in enzyme activity. Superpositions between the PfSpdS enzyme–inhibitor complex and apo-KlSpdS structures (Figure 4E,F) revealed that the residues of PfSpdS did not align well with those of KlSpdS in SpdS–adoDATO complexes compared to HsSpdS. However, all residues in SpdS–4MCHA complex corresponded for HsSpdS as well as PfSpdS. Taken together, these findings suggest that SpdS exhibits different inhibitory effects when complexed with adoDATO and 4MCHA.

3. Discussion

Spermidine is produced from putrescine by SpdS. Although the molecular mechanism underlying SpdS function is well-established, its structure-based evolutionary relationships remain to be fully understood; moreover, very few structural studies have been conducted on SpdS from fungal species. Here, we found that the first structure of fungal KlSpdS, which was determined at 1.9 Å resolution, exhibited highly similar to that of HsSpdS, suggesting that KlSpdS is phylogenetically closer to HsSpdS than SpdS from E. coli and could therefore utilize a sequential mechanism rather than a ping-pong mechanism [19,20,21,22,23]. In KlSpdS, the carboxylate group of Asp167 plays a major role in substrate deprotonation along with the aid of the backbone carbonyl of Ser168 and hydroxyl groups of conserved residues Tyr73 and Tyr235 [1,29]. Meanwhile, the carboxylate group of Asp170 plays an essential role in putrescine binding by anchoring the end of the diamine [1,27], whereas the carboxylate group of Asp98 is involved in binding the N1 atom of spermidine to the aminopropyl group of dcSAM [1,26]. Asp98 also promotes the initiation of nucleophilic attack on dcSAM by anchoring the aminopropyl group and fixing it in an appropriate position to initiate the enzymatic reaction [1,22].
In general, the gate-keeping loop plays a crucial role in the enzymatic reaction of SpdS through three distinct modes of action [3,22,30]. Firstly, the loop covers the active site of SpdS. Superposition of complexed HsSpdS and apo-KlSpdS revealed that the gate-keeping loops have distinct conformations depending on ligand binding status. Secondly, gate-keeping loops are important for substrate recognition in SpdS. For instance, the substrate specificity of SpdS can be altered through site-directed mutations of the proline residue in the gate-keeping loop (corresponding to Pro174 in KlSpdS) of E. coli [25]. Finally, the gate-keeping loop stabilizes the active site by adopting a closed conformation. After the substrate binds to the active site, the conformation of the gate-keeping loop is changed through a series of enzymatic processes [1,24]. The gate-keeping loop was disordered in the apo structures of CeSpdS and TcSpdS but was well-ordered in the structures of PfSpdS complexed with adoDATO, dcSAM, and dcSAM–4MCHA [7,27,29].
Taken together, the analyses of the crystal structure of KlSpdS performed in this study provide insights into the structural diversity of SpdS. Despite these important findings, further studies are required to investigate two key aspects related to KlSpdS. First, structural and functional studies with various ligands are essential to reveal the reaction mechanism of KlSpdS. In addition, studies of SpdS from other fungal species are required to understand their molecular structure-based phylogenetic relationships with SpdS homologs.

4. Materials and Methods

4.1. Preparation of KlSpdS Expression Constructs

The gene encoding KlSpdS (NCBI ID: XP_451945) was amplified from K. lactis genomic DNA (Korean Collection for Type Cultures, Daejeon, Republic of Korea) using polymerase chain reaction (PCR), as described previously [31]. All amplified fragments were digested using NdeI and XhoI restriction enzymes (R006S and R007S, respectively; Enzynomics, Republic of Korea) in a heating block at 37 °C for 4 h. The digested fragments were ligated with the pET28a and pET26b vectors using T4 ligase (M0202S; Roche, Germany) overnight at 18 °C to insert a hexahistidine (His6)-tag at either the N- or C-terminus of the target protein. The resulting vectors were subsequently transformed into the E. coli strain DH5α using kanamycin (AppliChem, Darmstadt, Germany) as a selection marker. The transformants were confirmed by colony PCR. All oligonucleotide primers used in this study were purchased from Cosmo Genetech (Seoul, Republic of Korea).

4.2. Purification of Recombinant Proteins

Plasmids encoding the KlSpdS protein were transformed into E. coli strain BL21 (DE3) Star. Cells were grown at 37 °C in Luria–Bertani medium (Ambrothia, Republic of Korea) containing 50 mg/L kanamycin (AppliChem) to an optical density at 600 nm (OD600) of approximately 0.6. Following induction with 0.3 mM isopropyl β-D-1-thiogalactopyranoside (IPTG; Calbiochem, Germany), the cells were further grown for 16 h at 20 °C, harvested by centrifugation at 3000 rpm at 4 °C for 20 min, and resuspended in a buffer containing 20 mM Tris (pH 8.0; Sigma–Aldrich, St. Louis, MO, USA), 250 mM NaCl (AppliChem), 5% glycerol (Affymetrix, Santa Clara, CA, USA), 0.2% Triton X-100 (Sigma–Aldrich), 10 mM β-mercaptoethanol (BioBasic, Markham, ON, Canada), and 0.2 mM phenylmethylsulfonyl fluoride (Sigma–Aldrich). Next, cells were disrupted by ultrasonication (VCX-500/750, Sonics, Newtown, CT, USA) with 3-s pulse-on and 3-s pulse-off cycles continuously for 15 min. Cell debris was removed by centrifugation at 13,000 rpm for 40 min, and the supernatant was bound to Ni–NTA agarose (Qiagen, Hilden, Germany) at 7 °C for 90 min. After washing with His-binding buffer (300 mM NaCl, 50 mM Tris, pH 8.0) containing 5 mM imidazole (Sigma–Aldrich), bound proteins were eluted with His-elution buffer (200 mM NaCl, 50 mM Tris, pH 8.0) containing 250 mM imidazole (Sigma–Aldrich). Purified proteins were subjected to size-exclusion chromatography (SEC) using a HiPrep 16/60 Sephacryl S-300 HR column (GE Healthcare, Chicago, IL, USA) and an eluent buffer containing 20 mM Tris (pH 7.5), 150 mM NaCl, and 2 mM dithiothreitol (DTT; Calbiochem). Following SEC, proteins were stored at −80 °C until crystallization. Protein purity was assessed by performing sodium dodecyl sulfate-polyacrylamide gel electrophoresis using a 15% acrylamide gel, which produced a single band corresponding to the calculated molecular weight of the target protein.

4.3. Crystallization and Improvements

All crystallization experiments were performed at 20 °C using the sitting-drop vapor diffusion method in 96-well sitting-drop plates (Art Robbins Instruments, Sunnyvale, CA, USA). Approximately 600 different conditions from sparse-matrix screening solution kits were tested to identify the optimal crystallization conditions. The following kits were used: PEG/Ion (HR2-126 and −098), Index (HR2-144), Salt Rx 1/2 (HR2-107 and -109), and Crystal Screen 1/2 (HR2-110 and -112) from Hampton Research (Viejo, CA, USA), Wizard 1/2 (CS-311, Jena Bioscience, Germany), and SG1 Screen (MD1-88, Molecular Dimensions, Rotherham, UK). KlSpdS crystals grew within 24 h in drops containing equal volumes (1 μL) of protein sample (10 mg/mL in 150 mM NaCl, 2 mM DTT, and 20 mM Tris, pH 7.5) and reservoir solution (9.2% v/v TacsimateTM pH 5.0, 16.5% w/v PEG 3350). Additional screening was performed using additive (HR2-428, Hampton Research) and detergent (HR2-406, Hampton Research) screening kits. The optimal crystallization conditions used 9.2% v/v TacsimateTM (pH 5.0), 16.5% (w/v) PEG 3350, and 2.5% (v/v) 1-butanol.

4.4. Data Collection and Structure Determination

Prior to data collection, 30% glycerol was added to the reservoir solutions as a cryoprotectant, and crystals were flash-cooled in liquid nitrogen. All diffraction datasets were collected at 100 K on a beamline 5C at the Pohang Accelerator Laboratory (PAL, Republic of Korea) using a Quantum 270 CCD detector (USA). Data were processed using the HKL–2000 software suite (HKL Research, Charlottesville, VA, USA).
Experimental electron density maps were obtained by molecular replacement methods in Phenix software version 1.9 (Phenix Software International, Berkeley, CA, USA) and interpreted using the WinCoot program with Homo sapiens SpdS (HsSpdS; PBD code, 2O06) as a search model [32,33]. The details of data collection and the statistics used in this study are listed in Table 2.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/molecules28083446/s1.

Author Contributions

Conceptualization, J.H.C.; Validation, S.K.; Formal analysis, S.K.; Investigation, J.H.C.; Resources, J.H.C.; Data curation, S.K.; Writing—original draft, S.K.; Writing—review & editing, J.H.C.; Visualization, S.K.; Supervision, J.H.C.; Funding acquisition, J.H.C. All authors have read and agreed to the published version of the manuscript.

Funding

This research was supported by Basic Science Research Program through the National Research Foundation of Korea, funded by the Ministry of Science and ICT (grant No. NRF-2022R1F1A1073775 to JHC).

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We would like to thank beamline staff member Yeon-Gil Kim at beamline 5C of the Pohang Accelerator Laboratory (Pohang, Republic of Korea) for the data collection. We also thank Jung Min Choi for the helpful discussions.

Conflicts of Interest

The authors declare no conflict of interest.

Sample Availability

Samples of the compounds are not available from the authors.

References

  1. Wu, H.; Min, J.; Ikeguchi, Y.; Zeng, H.; Dong, A.; Loppnau, P.; Pegg, A.E.; Plotnikov, A.N. Structure and mechanism of spermidine synthases. Biochemistry 2007, 46, 8331–8339. [Google Scholar] [CrossRef] [PubMed]
  2. Miller-Fleming, L.; Olin-Sandoval, V.; Campbell, K.; Ralser, M. Remaining Mysteries of Molecular Biology: The Role of Polyamines in the Cell. J. Mol. Biol. 2015, 427, 3389–3406. [Google Scholar] [CrossRef] [PubMed]
  3. Zhou, X.; Chua, T.K.; Tkaczuk, K.L.; Bujnicki, J.M.; Sivaraman, J. The crystal structure of Escherichia coli spermidine synthase SpeE reveals a unique substrate-binding pocket. J. Struct. Biol. 2010, 169, 277–285. [Google Scholar] [CrossRef] [PubMed]
  4. Agostinelli, E.; Marques, M.P.; Calheiros, R.; Gil, F.P.; Tempera, G.; Viceconte, N.; Battaglia, V.; Grancara, S.; Toninello, A. Polyamines: Fundamental characters in chemistry and biology. Amino Acids. 2010, 38, 393–403. [Google Scholar] [CrossRef] [PubMed]
  5. Terui, Y.; Ohnuma, M.; Hiraga, K.; Kawashima, E.; Oshima, T. Stabilization of nucleic acids by unusual polyamines produced by an extreme thermophile, Thermus thermophilus. Biochem. J. 2005, 388, 427–433. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  6. Okada, K.; Hidese, R.; Fukuda, W.; Niitsu, M.; Takao, K.; Horai, Y.; Umezawa, N.; Higuchi, T.; Oshima, T.; Yoshikawa, Y.; et al. Identification of a novel aminopropyltransferase involved in the synthesis of branched-chain polyamines in hyperthermophiles. J. Bacteriol. 2014, 196, 1866–1876. [Google Scholar] [CrossRef] [Green Version]
  7. Dufe, V.T.; Luersen, K.; Eschbach, M.L.; Haider, N.; Karlberg, T.; Walter, R.D.; Al-Karadaghi, S. Cloning, expression, characterisation and three-dimensional structure determination of Caenorhabditis elegans spermidine synthase. FEBS Lett. 2005, 579, 6037–6043. [Google Scholar] [CrossRef] [Green Version]
  8. Gevrekci, A.O. The roles of polyamines in microorganisms. World J. Microbiol. Biotechnol. 2017, 33, 204. [Google Scholar] [CrossRef]
  9. Sagar, N.A.; Tarafdar, S.; Agarwal, S.; Tarafdar, A.; Sharma, S. Polyamines: Functions, Metabolism, and Role in Human Disease Management. Med. Sci. 2021, 9, 44. [Google Scholar] [CrossRef]
  10. Pietrocola, F.; Lachkar, S.; Enot, D.P.; Niso-Santano, M.; Bravo-San Pedro, J.M.; Sica, V.; Izzo, V.; Maiuri, M.C.; Madeo, F.; Marino, G.; et al. Spermidine induces autophagy by inhibiting the acetyltransferase EP300. Cell Death Differ. 2015, 22, 509–516. [Google Scholar] [CrossRef]
  11. Saini, P.; Eyler, D.E.; Green, R.; Dever, T.E. Hypusine-containing protein eIF5A promotes translation elongation. Nature 2009, 459, 118–121. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  12. Park, M.H.; Wolff, E.C. Hypusine, a polyamine-derived amino acid critical for eukaryotic translation. J. Biol. Chem. 2018, 293, 18710–18718. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Schubert, H.L.; Blumenthal, R.M.; Cheng, X. Many paths to methyltransfer: A chronicle of convergence. Trends Biochem. Sci. 2003, 28, 329–335. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Kozbial, P.Z.; Mushegian, A.R. Natural history of S-adenosylmethionine-binding proteins. BMC Struct. Biol. 2005, 5, 19. [Google Scholar] [CrossRef] [Green Version]
  15. Kimura, S.; Miyauchi, K.; Ikeuchi, Y.; Thiaville, P.C.; Crecy-Lagard, V.; Suzuki, T. Discovery of the beta-barrel-type RNA methyltransferase responsible for N6-methylation of N6-threonylcarbamoyladenosine in tRNAs. Nucleic. Acids Res. 2014, 42, 9350–9365. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  16. Fontecave, M.; Atta, M.; Mulliez, E. S-adenosylmethionine: Nothing goes to waste. Trends Biochem. Sci. 2004, 29, 243–249. [Google Scholar] [CrossRef]
  17. Tabor, C.W.; Tabor, H. Polyamines. Annu. Rev. Biochem. 1984, 53, 749–790. [Google Scholar] [CrossRef]
  18. Pegg, A.E.; McCann, P.P. Polyamine metabolism and function. Am. J. Physiol. 1982, 243, C212–C221. [Google Scholar] [CrossRef] [Green Version]
  19. Yoon, S.O.; Lee, Y.S.; Lee, S.H.; Cho, Y.D. Polyamine synthesis in plants: Isolation and characterization of spermidine synthase from soybean (Glycine max) axes. Biochim. Biophys. Acta 2000, 1475, 17–26. [Google Scholar] [CrossRef]
  20. Zappia, V.; Cacciapuoti, G.; Pontoni, G.; Oliva, A. Mechanism of propylamine-transfer reactions. Kinetic and inhibition studies on spermidine synthase from Escherichia coli. J. Biol. Chem. 1980, 255, 7276–7280. [Google Scholar] [CrossRef]
  21. Haider, N.; Eschbach, M.L.; Dias Sde, S.; Gilberger, T.W.; Walter, R.D.; Luersen, K. The spermidine synthase of the malaria parasite Plasmodium falciparum: Molecular and biochemical characterisation of the polyamine synthesis enzyme. Mol. Biochem. Parasitol. 2005, 142, 224–236. [Google Scholar] [CrossRef] [PubMed]
  22. Korolev, S.; Ikeguchi, Y.; Skarina, T.; Beasley, S.; Arrowsmith, C.; Edwards, A.; Joachimiak, A.; Pegg, A.E.; Savchenko, A. The crystal structure of spermidine synthase with a multisubstrate adduct inhibitor. Nat. Struct. Biol. 2002, 9, 27–31. [Google Scholar] [CrossRef]
  23. Samejima, K.; Yamanoha, B. Purification of spermidine synthase from rat ventral prostate by affinity chromatography on immobilized S-adenosyl(5′)-3-thiopropylamine. Arch. Biochem. Biophys. 1982, 216, 213–222. [Google Scholar] [CrossRef]
  24. Guedez, G.; Pothipongsa, A.; Siren, S.; Liljeblad, A.; Jantaro, S.; Incharoensakdi, A.; Salminen, T.A. Crystal structure of dimeric Synechococcus spermidine synthase with bound polyamine substrate and product. Biochem. J. 2019, 476, 1009–1020. [Google Scholar] [CrossRef] [PubMed]
  25. Lee, M.J.; Yang, Y.T.; Lin, V.; Huang, H. Site-directed mutations of the gatekeeping loop region affect the activity of Escherichia coli spermidine synthase. Mol. Biotechnol. 2013, 54, 572–580. [Google Scholar] [CrossRef] [PubMed]
  26. Sekula, B.; Dauter, Z. Spermidine Synthase (SPDS) Undergoes Concerted Structural Rearrangements Upon Ligand Binding—A Case Study of the Two SPDS Isoforms From Arabidopsis thaliana. Front. Plant Sci. 2019, 10, 555. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  27. Dufe, V.T.; Qiu, W.; Muller, I.B.; Hui, R.; Walter, R.D.; Al-Karadaghi, S. Crystal structure of Plasmodium falciparum spermidine synthase in complex with the substrate decarboxylated S-adenosylmethionine and the potent inhibitors 4MCHA and AdoDATO. J. Mol. Biol. 2007, 373, 167–177. [Google Scholar] [CrossRef]
  28. Lu, P.K.; Tsai, J.Y.; Chien, H.Y.; Huang, H.; Chu, C.H.; Sun, Y.J. Crystal structure of Helicobacter pylori spermidine synthase: A Rossmann-like fold with a distinct active site. Proteins 2007, 67, 743–754. [Google Scholar] [CrossRef]
  29. Amano, Y.; Namatame, I.; Tateishi, Y.; Honboh, K.; Tanabe, E.; Niimi, T.; Sakashita, H. Structural insights into the novel inhibition mechanism of Trypanosoma cruzi spermidine synthase. Acta Crystallogr. D Biol. Crystallogr. 2015, 71, 1879–1889. [Google Scholar] [CrossRef]
  30. Seckute, J.; McCloskey, D.E.; Thomas, H.J.; Secrist, J.A., III; Pegg, A.E.; Ealick, S.E. Binding and inhibition of human spermidine synthase by decarboxylated S-adenosylhomocysteine. Protein. Sci. 2011, 20, 1836–1844. [Google Scholar] [CrossRef]
  31. Kim, S.; Nguyen, G.T.; Chang, J.H. Purification, crystallization, and X-ray crystallographic analysis of spermidine synthase from Kluyveromyces lactis. BioDesign 2021, 9, 36–40. [Google Scholar] [CrossRef]
  32. Emsley, P.; Lohkamp, B.; Scott, W.G.; Cowtan, K. Features and development of Coot. Acta Crystallogr. D Biol. Crystallogr. 2010, 66, 486–501. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Liebschner, D.; Afonine, P.V.; Baker, M.L.; Bunkoczi, G.; Chen, V.B.; Croll, T.I.; Hintze, B.; Hung, L.W.; Jain, S.; McCoy, A.J.; et al. Macromolecular structure determination using X-rays, neutrons and electrons: Recent developments in Phenix. Acta Crystallogr. D Struct. Biol. 2019, 75, 861–877. [Google Scholar] [CrossRef] [PubMed] [Green Version]
Figure 1. Overall structure of spermidine synthase (SpdS) from Kluyveromyces lactis (KlSpdS). (A) The dimeric structure of KlSpdS is shown as a cartoon. Chain A is shown in lime green, and chain B is shown in pink. Different view of 90° rotation along the x-axis is shown in right panel. (B) Monomeric structure of chain B in KlSpdS. The N-terminal domain is colored green. The catalytic core domain is shown in orange. The C-terminal domain is shown in cyan. The active site of KlSpdS is highlighted in the red-dashed circle. The gate-keeping loop is highlighted in the blue-dashed circle. (C) Electrostatic surface model of the KlSpdS monomer. Red and blue represent negatively and positively charged surfaces, respectively. The active site of KlSpdS is highlighted in the yellow-dashed circle.
Figure 1. Overall structure of spermidine synthase (SpdS) from Kluyveromyces lactis (KlSpdS). (A) The dimeric structure of KlSpdS is shown as a cartoon. Chain A is shown in lime green, and chain B is shown in pink. Different view of 90° rotation along the x-axis is shown in right panel. (B) Monomeric structure of chain B in KlSpdS. The N-terminal domain is colored green. The catalytic core domain is shown in orange. The C-terminal domain is shown in cyan. The active site of KlSpdS is highlighted in the red-dashed circle. The gate-keeping loop is highlighted in the blue-dashed circle. (C) Electrostatic surface model of the KlSpdS monomer. Red and blue represent negatively and positively charged surfaces, respectively. The active site of KlSpdS is highlighted in the yellow-dashed circle.
Molecules 28 03446 g001
Figure 2. Comparisons of the gate-keeping loop between the apo structure of spermidine synthase from Kluyveromyces lactis (KlSpdS) and four other species. Each figure represents the superposition of the monomers to compare the gate-keeping loop. In (AD), the monomer of the apo-KlSpdS is colored orange. (A) Superposition of apo-KlSpdS with that of SpdS from Homo sapiens (HsSpdS; PDB code 2O0L). The HsSpdS monomer is shown in light blue. (B) Superposition of the apo-KlSpdS with that of SpdS from Arabidopsis thaliana (AtSpdS; PDB code 6O63). The AtSpdS monomer is shown in yellow. (C) Superposition of apo-KlSpdS with that of SpdS from Thermus thermophilus (TtSpdS; PDB code 1UIR). The TtSpdS monomer is shown in cyan. (D) Superposition of apo-KlSpdS with that of SpdS from Thermotoga maritima (TmSpdS; PDB code 1INL). The TmSpdS monomer is shown in lime green. The purple color indicates the gate-keeping loop of KlSpdS, wherease, the gate-keeping loops from compared structures were colored in blue.
Figure 2. Comparisons of the gate-keeping loop between the apo structure of spermidine synthase from Kluyveromyces lactis (KlSpdS) and four other species. Each figure represents the superposition of the monomers to compare the gate-keeping loop. In (AD), the monomer of the apo-KlSpdS is colored orange. (A) Superposition of apo-KlSpdS with that of SpdS from Homo sapiens (HsSpdS; PDB code 2O0L). The HsSpdS monomer is shown in light blue. (B) Superposition of the apo-KlSpdS with that of SpdS from Arabidopsis thaliana (AtSpdS; PDB code 6O63). The AtSpdS monomer is shown in yellow. (C) Superposition of apo-KlSpdS with that of SpdS from Thermus thermophilus (TtSpdS; PDB code 1UIR). The TtSpdS monomer is shown in cyan. (D) Superposition of apo-KlSpdS with that of SpdS from Thermotoga maritima (TmSpdS; PDB code 1INL). The TmSpdS monomer is shown in lime green. The purple color indicates the gate-keeping loop of KlSpdS, wherease, the gate-keeping loops from compared structures were colored in blue.
Molecules 28 03446 g002
Figure 3. Structural comparisons of the three key aspartic residues in the active site of spermidine synthase (SpdS) from five different species. In (AD), the Kluyveromyces lactis SpdS (KlSpdS) monomer is shown in orange. (A) The Homo sapiens SpdS (HsSpdS) monomer is shown in light blue. (B) The Arabidopsis thaliana SpdS (AtSpdS) monomer is shown in warm pink. (C) The Plasmodium falciparum SpdS (PfSpdS) monomer is shown in lime green. (D) The Trypanosoma cruzi SpdS (TcSpdS) monomer is shown in yellow.
Figure 3. Structural comparisons of the three key aspartic residues in the active site of spermidine synthase (SpdS) from five different species. In (AD), the Kluyveromyces lactis SpdS (KlSpdS) monomer is shown in orange. (A) The Homo sapiens SpdS (HsSpdS) monomer is shown in light blue. (B) The Arabidopsis thaliana SpdS (AtSpdS) monomer is shown in warm pink. (C) The Plasmodium falciparum SpdS (PfSpdS) monomer is shown in lime green. (D) The Trypanosoma cruzi SpdS (TcSpdS) monomer is shown in yellow.
Molecules 28 03446 g003
Figure 4. Superposition of the KlSpdS structure with SpdS-ligand complexes. (A) Superposition of the structures of spermidine synthase (SpdS) from Kluyveromyces lactis (KlSpdS) and the Homo sapiens (HsSpdS)–putrescine complex. The substrate putrescine is shown in yellow. The panel provides a detailed view of the interaction between putrescine and KlSpdS, with overlaid HsSpdS. KlSpdS and HsSpdS residues are shown in orange and light blue, respectively. (B) Superposition of the structures of KlSpdS and the HsSpdS–spermidine complex. The product spermidine is shown in green. The panel provides a detailed view of the interaction between spermidine and KlSpdS, with overlaid HsSpdS. KlSpdS and HsSpdS residues are shown in orange and light blue, respectively. (C) Superposition of KlSpdS and the HsSpdS–dcSAM complex structures. The cofactor dcSAM is shown in magenta. The panel provides a detailed view of the interaction between dcSAM and KlSpdS, with overlaid HsSpdS. KlSpdS and HsSpdS residues are shown in orange and light blue, respectively. Abbreviations: dcSAM, decarboxylated S-adenosylmethionine. (D) Superposition of KlSpdS and the HsSpdS–MTA complex structures. The byproduct MTA is shown in cyan. The panel provides a detailed view of the interaction between MTA and KlSpdS, with overlaid HsSpdS. KlSpdS and HsSpdS residues are shown in orange and light blue, respectively. Abbreviations: MTA, methylthio-adenosine. (E) Superposition of KlSpdS and the PfSpdS–adoDATO complex structures. The inhibitor adoDATO is shown in yellow. The panel provides a detailed view of the interaction between adoDATO and KlSpdS, with overlaid PfSpdS. KlSpdS and PfSpdS residues are shown in orange and lime green, respectively. Abbreviations: adoDATO, S-adenosyl-1,8-diamino-3-thiooctane. (F) Superposition of KlSpdS and the PfSpdS–4MCHA complex structures. The inhibitor 4MCHA is shown in magenta. The panel provides a detailed view of the interaction between 4MCHA and KlSpdS, with overlaid PfSpdS. KlSpdS and PfSpdS residues are shown in orange and lime green, respectively. Abbreviations: 4MCHA, trans-4-methylcyclohexylamine.
Figure 4. Superposition of the KlSpdS structure with SpdS-ligand complexes. (A) Superposition of the structures of spermidine synthase (SpdS) from Kluyveromyces lactis (KlSpdS) and the Homo sapiens (HsSpdS)–putrescine complex. The substrate putrescine is shown in yellow. The panel provides a detailed view of the interaction between putrescine and KlSpdS, with overlaid HsSpdS. KlSpdS and HsSpdS residues are shown in orange and light blue, respectively. (B) Superposition of the structures of KlSpdS and the HsSpdS–spermidine complex. The product spermidine is shown in green. The panel provides a detailed view of the interaction between spermidine and KlSpdS, with overlaid HsSpdS. KlSpdS and HsSpdS residues are shown in orange and light blue, respectively. (C) Superposition of KlSpdS and the HsSpdS–dcSAM complex structures. The cofactor dcSAM is shown in magenta. The panel provides a detailed view of the interaction between dcSAM and KlSpdS, with overlaid HsSpdS. KlSpdS and HsSpdS residues are shown in orange and light blue, respectively. Abbreviations: dcSAM, decarboxylated S-adenosylmethionine. (D) Superposition of KlSpdS and the HsSpdS–MTA complex structures. The byproduct MTA is shown in cyan. The panel provides a detailed view of the interaction between MTA and KlSpdS, with overlaid HsSpdS. KlSpdS and HsSpdS residues are shown in orange and light blue, respectively. Abbreviations: MTA, methylthio-adenosine. (E) Superposition of KlSpdS and the PfSpdS–adoDATO complex structures. The inhibitor adoDATO is shown in yellow. The panel provides a detailed view of the interaction between adoDATO and KlSpdS, with overlaid PfSpdS. KlSpdS and PfSpdS residues are shown in orange and lime green, respectively. Abbreviations: adoDATO, S-adenosyl-1,8-diamino-3-thiooctane. (F) Superposition of KlSpdS and the PfSpdS–4MCHA complex structures. The inhibitor 4MCHA is shown in magenta. The panel provides a detailed view of the interaction between 4MCHA and KlSpdS, with overlaid PfSpdS. KlSpdS and PfSpdS residues are shown in orange and lime green, respectively. Abbreviations: 4MCHA, trans-4-methylcyclohexylamine.
Molecules 28 03446 g004
Table 1. Structural similarity comparison for homologous structures of spermidine synthase among five species using Dali a.
Table 1. Structural similarity comparison for homologous structures of spermidine synthase among five species using Dali a.
SpeciesZ-ScoreRMS
Deviation (Å)
Identity
(%)
CαPDB
Code
Homo sapiens43.30.9572892O06
Plasmodium falciparum40.61.4482702HTE
Arabidopsis thaliana40.41.4492851XJ5
Caenorhabditis elegans40.11.1562762B2C
Trypanosoma cruzi40.01.5442944YUV
a This server computes optimal and suboptimal structural alignments between two protein structures using the DaliLite-pairwise option. Available online: http://ekhidna.biocenter.helsinki.fi/dali/ (accessed on 5 September 2019).
Table 2. Data collection and refinement statics for KlSpdS.
Table 2. Data collection and refinement statics for KlSpdS.
StatisticsKlSpdS
Data collection
Space groupP212121
Cell dimensions (Å)
a, b, c (Å)65.252, 98.180, 102.134
α, β, γ (°)90, 90, 90
Resolution range (Å) a50.0–1.9
No. of reflections676,985
No. of unique reflections52,396
Rmerge b (%)13.5 (45.2)
I/σ (I)33.0 (4.9)
Completeness (%)100 (100)
Redundancy12.9 (12.1)
CC1/20.996 (0.965)
Structure refinement
Resolution (Å)48.0–1.9
No. of reflections676,985 (52,396)
Rwork c/Rfree16.8/19.9
No. atoms
Protein4639
Water555
R.m.s. deviation
Bond lengths (Å)0.007
Angles (°)0.818
Average B-factor (Å2)23.7
Ramachandran plot (%)
Favored region97.4
Outliers0.0
PDB code8IYI
a Numbers in parentheses are statistics from the highest resolution shell. b Rmerge = Σ |IobsIavg|/Iobs, where Iobs is the observed intensity of individual reflections and Iavg is the average over symmetry equivalents. c Rwork = Σ ||Fo| − |Fc||/Σ |Fo|, where |Fo| and |Fc| are the observed and calculated structure factor amplitudes, respectively. Rfree was calculated using 5% of the data.
Disclaimer/Publisher’s Note: The statements, opinions and data contained in all publications are solely those of the individual author(s) and contributor(s) and not of MDPI and/or the editor(s). MDPI and/or the editor(s) disclaim responsibility for any injury to people or property resulting from any ideas, methods, instructions or products referred to in the content.

Share and Cite

MDPI and ACS Style

Kim, S.; Chang, J.H. Structural Analysis of Spermidine Synthase from Kluyveromyces lactis. Molecules 2023, 28, 3446. https://doi.org/10.3390/molecules28083446

AMA Style

Kim S, Chang JH. Structural Analysis of Spermidine Synthase from Kluyveromyces lactis. Molecules. 2023; 28(8):3446. https://doi.org/10.3390/molecules28083446

Chicago/Turabian Style

Kim, Seongjin, and Jeong Ho Chang. 2023. "Structural Analysis of Spermidine Synthase from Kluyveromyces lactis" Molecules 28, no. 8: 3446. https://doi.org/10.3390/molecules28083446

Article Metrics

Back to TopTop