Design, expression, purification and crystallization of human 14-3-3ζ protein chimera with phosphopeptide from proapoptotic protein BAD

https://doi.org/10.1016/j.pep.2020.105707Get rights and content

Highlights

  • 14-3-3ζ protein chimera with the BAD peptide is designed, purified and crystallized.

  • BAD peptide SRHSS75YPAGTE is phosphorylated upon co-expression with PKA in E.coli.

  • Tethered phosphorylated BAD peptide occupies the amphipathic grooves of 14-3-3ζ.

  • Efficient purification of the untagged 14-3-3ζ-BAD75 chimera is demonstrated.

  • The assembled chimera readily crystallizes in different chemical environments.

Abstract

14-3-3 protein isoforms regulate multiple processes in eukaryotes, including apoptosis and cell division. 14-3-3 proteins preferentially recognize phosphorylated unstructured motifs, justifying the protein-peptide binding approach to study 14-3-3/phosphotarget complexes. Tethering of human 14-3-3σ with partner phosphopeptides via a short linker has provided structural information equivalent to the use of synthetic phosphopeptides, simultaneously facilitating purification and crystallization. Nevertheless, the broader applicability to other 14-3-3 isoforms and phosphopeptides was unclear. Here, we designed a novel 14-3-3ζ chimera with a conserved phosphopeptide from BAD, whose complex with 14-3-3 is a gatekeeper of apoptosis regulation. The chimera could be bacterially expressed and purified without affinity tags. Co-expressed PKA efficiently phosphorylates BAD within the chimera and blocks its interaction with a known 14-3-3 phosphotarget, suggesting occupation of the 14-3-3 grooves by the tethered BAD phosphopeptide. Efficient crystallization of the engineered protein suggests suitability of the “chimeric” approach for studies of other relevant 14-3-3 complexes.

Introduction

The eukaryotic 14-3-3 protein family comprises ~30-kDa polypeptides that assemble into homo- and heterodimers performing various roles in the regulation of apoptosis, cell division, transcription, signal transduction and other important cellular processes [1,2]. Such rich functionality is achieved due to rather promiscuous binding of 14-3-3 to hundreds of target proteins with phosphorylated serine or threonine in 14-3-3 binding motifs [1,3,4]. These motifs tend to be located in the intrinsically disordered regions, which facilitate accommodation of the phosphopeptides in the amphipathic grooves (AGs) of 14-3-3 [5,6]. Although pioneering studies outlined binding preferences of certain phosphorylated peptide sequences, revealing at least three canonical 14-3-3 binding motifs (I-III) [4,7], the constantly expanding set of experimentally confirmed 14-3-3 binding sequences helped realize that those sequence determinants are not as strict as previously thought. Indeed, there are now many confirmed 14-3-3 binders that lack the canonical binding sequences, which complicates the prediction of 14-3-3 binding sites [8]. As the interactome of 14-3-3 proteins is estimated to exceed 1,000 proteins [6,9], and some of them may be polyvalent 14-3-3 binders [10], there is urgent need to systematically analyze as many as possible 14-3-3-bound phosphopeptide conformations.

Not surprisingly, the wide interactome implicates 14-3-3 proteins in the development of many human disorders, including neurodegenerative, developmental and oncological abnormalities. In this sense, the modulation of the 14-3-3 interaction with phosphorylated target proteins is recognized as a big opportunity for drug discovery and therapy [11,12]. Toward this goal, 14-3-3/phosphotarget complexes must be well characterized structurally. Yet the 14-3-3 interactome remains disproportionally explored [8] – among hundreds of experimentally confirmed 14-3-3 protein-protein interactions (PPIs) [13], only a minority has been structurally characterized with phosphopeptides, and even fewer have been with full protein partners (e.g., AANAT [14], HSPB6 [15], Hd3a [16], Nth1 [17], ExoS/ExoT [18] and B-RAF [19]).

Several 14-3-3 isoforms are usually expressed per organism and there are seven 14-3-3 isoforms in humans (β, ζ, γ, ε, σ, τ, η) [20]. Although sharing a conserved and unique α-helical fold [21,22], various 14-3-3 isoforms reveal distinct preferences towards dimerization with other isoforms and may also display some specialization [6,20,23]. For example, 14-3-3σ almost exclusively homodimerizes, whereas 14-3-3ε prefers to heterodimerize with other 14-3-3 isoforms [21,24,25]. The rest isoforms are intermediate in this ability [21,25]. Likewise, 14-3-3 isoforms appear to differ in terms of their regulation by posttranslational modifications [26,27]. For example, phosphorylation of a semi-conserved residue Ser58 located in the dimer interface by a range of protein kinases including PKA was shown to promote 14-3-3 monomerization [[28], [29], [30], [31], [32]], which in turn may interfere with the phosphopeptide-binding ability directly [33] or indirectly [34]. Ser58 is replaced by alanine in 14-3-3σ and 14-3-3τ, which makes them unphosphorylatable at this position.

The remarkable structural element of 14-3-3 proteins is their intrinsically disordered C-terminal tail of 12–23 residues. This element is the most variable segment in 14-3-3 proteins. One of its functions in 14-3-3ζ was proposed to be an autoinhibitory role based on its transient interaction with the AGs of 14-3-3, potentially excluding undesired binding of weak and unspecific ligands [22,35]. Curiously, the C-terminal segment of Cryptosporidium parvum 14-3-3 was found spontaneously phosphorylated and bound in the AG (PDB ID 3EFZ [36]) (Fig. 1A). Later on, this serendipitous discovery inspired the design of the first artificial chimera of human 14-3-3σ with phosphorylated fragments from partner proteins, to streamline purification and crystallization of 14-3-3/phosphopeptide complexes at a fixed protein-peptide stoichiometry. These advantages remove many challenges associated with traditional peptide co-crystallizations, such as low peptide solubility, weak binding affinity and incomplete occupancy [37]. In addition, the 3EFZ crystal structure provided information on the length of the linker, because all residues of the phosphorylated C-terminal segment of C. parvum 14-3-3 were crystallographically resolved, with no apparent constraints for the phosphopeptide binding in the AG [36] (Fig. 1A). The prototype 14-3-3 chimera was successfully used to determine crystal structures of at least six 14-3-3/phosphopeptide complexes [37,38]. However, it was based on 14-3-3σ only and the question remained whether studies on other 14-3-3 isoforms can benefit from this approach, which would open up access to all the 14-3-3 isoform-specific client interactions.

Here, we expanded this approach by designing a chimera of a heterodimerizable 14-3-3 isoform, 14-3-3ζ. As a partner peptide we have chosen the conserved Ser75 (human BAD numbering, equivalent to Ser112 in mouse) phosphopeptide of the death agonist protein BAD. BAD is a well-known 14-3-3 partner phosphorylatable at several positions by PKA [39], PAK1 [40], PKB/Akt [41,42] and RSK1 [43] kinases. Phosphorylated BAD is sequestered by 14-3-3 in the cytoplasm until pro-apoptotic stimuli dissociate 14-3-3 from BAD, releasing it for translocation to the mitochondrial membrane, association with other members of the BCL-2 family and triggering of apoptosis [44,45]. The 14-3-3/phospho-BAD complex thus represents a gatekeeper of the pro- and anti-apoptotic regulation [[46], [47], [48], [49], [50], [51], [52], [53]] and an attractive target of drug discovery and therapy to modulate the balance in apoptotic and pro-survival signaling in a broad spectrum of pathological conditions [[54], [55], [56], [57]]. Intriguingly, despite its huge significance for biomedicine, the structure of the 14-3-3/BAD complex has not been resolved. As the first step in this direction, here we report on successful design, bacterial expression, untagged purification and efficient crystallization of the new 14-3-3ζ-BAD75 chimera and discuss further improvements and application of 14-3-3 chimeras in structural studies.

Section snippets

Cloning, protein expression and purification

A gene encoding a modified human 14-3-3ζΔC (Uniprot ID P63104; residues 1–229, S58A, 157KKE159→AAA) C-terminally fused with a GSGG linker to residues 71–81 of BAD (human numbering) was codon optimized for expression in Escherichia coli using Codon Optimization OnLine (COOL) [58] and synthesized by IDT Technologies (Coralville, Iowa, USA). The 14-3-3ζΔC gene was flanked by NdeI and BamHI restriction sites to permit convenient alteration of the 14-3-3 or BAD sequences. No affinity purification

Design of the 14-3-3ζ chimera with the selected BAD phosphopeptide

We have recently demonstrated the usefulness of 14-3-3 chimeras with phosphorylated fragments of partner proteins [37] for obtaining and studying 14-3-3 complexes by structural methods. Although convenient and promising, the approach has so far been tested on a small set of phosphopeptides and on human 14-3-3σ isoform exclusively [33], leaving the question of its applicability to other 14-3-3 isoforms and other phosphopeptides unaddressed. Answering this is important because not all

Discussion

We have recently shown that 14-3-3σ protein fusion with its cognate phosphopeptides can be used to facilitate production and structural studies of 14-3-3 complexes, whereas the appropriate linker length allows for the correct placement of the tethered phosphopeptide in the amphipathic binding grooves of 14-3-3 [37]. Given the remarkable isoform-specific roles of 14-3-3 in physiologically relevant processes [6,23,[72], [73], [74], [75]], the unanswered question of whether the proposed chimeric

Author contributions

K.V.T. expressed, purified and characterized proteins, A.R. and I.G. performed protein crystallization, R.B.C. and N.N.S. discussed the main idea and analyzed data, R.B.C. performed cloning, N.N.S. purified and characterized proteins, interpreted data and wrote the paper with input from K.V.T., I.G. and R.B.C.

Credit author statement

Kristina V. Tugaeva: Investigation, Methodology, Writing - Review & Editing. Alina Remeeva: Resources, Investigation. Ivan Gushchin: Resources, Investigation. Richard B. Cooley: Resources, Writing - Review & Editing. Nikolai N. Sluchanko: Conceptualization, Project administration, Methodology, Writing - Original Draft, Writing - Review & Editing, Visualization, Supervision, Funding acquisition

Declaration of competing interest

The authors declare no conflict of interest.

Acknowledgments

We are thankful to Daria I. Kalacheva for help with protein purification and to Ilia Chetviorkin for writing script to convert diode-array detector raw data. This work was supported by the Russian Science Foundation (grant no. 19-74-10031). R.B.C. acknowledges support from the Medical Research Foundation at Oregon Health Sciences University, and the Collins Medical Trust. Robotic crystallization (A.R. and I.G.) was supported by the Ministry of Science and Higher Education of the Russian

References (81)

  • E. Wilker et al.

    A structural basis for 14-3-3sigma functional specificity

    J. Biol. Chem.

    (2005)
  • A. Aitken

    Post-translational modification of 14-3-3 isoforms and regulation of cellular function

    Semin. Cell Dev. Biol.

    (2011)
  • Y.-M. Gu et al.

    Protein kinase A phosphorylates and regulates dimerization of 14-3-3 epsilon

    FEBS Lett.

    (2006)
  • N.N. Sluchanko et al.

    Effect of mutations mimicking phosphorylation on the structure and properties of human 14-3-3zeta

    Arch. Biochem. Biophys.

    (2008)
  • J.M. Woodcock et al.

    The dimeric versus monomeric status of 14-3-3zeta is controlled by phosphorylation of Ser58 at the dimer interface

    J. Biol. Chem.

    (2003)
  • D. Powell et al.

    Identification of 14-3-3zeta as a protein kinase B/Akt substrate

    J. Biol. Chem.

    (2002)
  • T. Gokirmak et al.

    Phosphomimetic mutation of a conserved serine residue in Arabidopsis thaliana 14-3-3omega suggests a regulatory role of phosphorylation in dimerization and target interactions

    Plant Physiol. Biochem.

    (2015)
  • N.N. Sluchanko et al.

    Phosphomimicking mutations of human 14-3-3zeta affect its interaction with tau protein and small heat shock protein HspB6

    Arch. Biochem. Biophys.

    (2011)
  • J. Silhan et al.

    14-3-3 protein C-terminal stretch occupies ligand binding groove and is displaced by phosphopeptide binding

    J. Biol. Chem.

    (2004)
  • H. Harada et al.

    Phosphorylation and inactivation of BAD by mitochondria-anchored protein kinase A

    Mol. Cell

    (1999)
  • S.R. Datta et al.

    Akt phosphorylation of BAD couples survival signals to the cell-intrinsic death machinery

    Cell

    (1997)
  • Y. Tan et al.

    p90(RSK) blocks bad-mediated cell death via a protein kinase C-dependent pathway

    J. Biol. Chem.

    (1999)
  • Y. Tan et al.

    BAD Ser-155 phosphorylation regulates BAD/Bcl-XL interaction and cell survival

    J. Biol. Chem.

    (2000)
  • J. Fan et al.

    A model of ischemia and reperfusion increases JNK activity, inhibits the association of BAD and 14-3-3, and induces apoptosis of rabbit spinal neurocytes

    Neurosci. Lett.

    (2010)
  • N.L. Bui et al.

    Bad phosphorylation as a target of inhibition in oncology

    Canc. Lett.

    (2018)
  • S.C. Masters et al.

    14-3-3 proteins mediate an essential anti-apoptotic signal

    J. Biol. Chem.

    (2001)
  • N.N. Sluchanko et al.

    Hidden disorder propensity of the N-terminal segment of universal adapter protein 14-3-3 is manifested in its monomeric form: novel insights into protein dimerization and multifunctionality

    Biochim. Biophys. Acta

    (2015)
  • Y. Aghazadeh et al.

    Hormone-induced 14-3-3gamma adaptor protein regulates steroidogenic acute regulatory protein activity and steroid biosynthesis in MA-10 Leydig cells

    J. Biol. Chem.

    (2012)
  • L.R. Jagemann et al.

    The functional interaction of 14-3-3 proteins with the ERK1/2 scaffold KSR1 occurs in an isoform-specific manner

    J. Biol. Chem.

    (2008)
  • Y.S. Kim et al.

    Protein kinase Cdelta is associated with 14-3-3 phosphorylation in seizure-induced neuronal death

    Epilepsy Res.

    (2010)
  • S.R. Datta et al.

    14-3-3 proteins and survival kinases cooperate to inactivate BAD by BH3 domain phosphorylation

    Mol. Cell

    (2000)
  • P. Astuti et al.

    Mitotic phosphorylation of Cdc25B Ser321 disrupts 14-3-3 binding to the high affinity Ser323 site

    J. Biol. Chem.

    (2010)
  • H. Fu et al.

    14-3-3 proteins: structure, function, and regulation

    Annu. Rev. Pharmacol. Toxicol.

    (2000)
  • D.M. Bustos et al.

    Intrinsic disorder is a key characteristic in partners that bind 14-3-3 proteins

    Proteins

    (2006)
  • F. Madeira et al.

    14-3-3-Pred: improved methods to predict 14-3-3-binding phosphopeptides

    Bioinformatics

    (2015)
  • N.N. Sluchanko

    Reading the phosphorylation code: binding of the 14-3-3 protein to multivalent client phosphoproteins

    Biochem. J.

    (2020)
  • L.M. Stevers et al.

    Modulators of 14-3-3 protein-protein interactions

    J. Med. Chem.

    (2018)
  • C. Johnson et al.

    Bioinformatic and experimental survey of 14-3-3-binding sites

    Biochem. J.

    (2010)
  • K. Taoka et al.

    14-3-3 proteins act as intracellular receptors for rice Hd3a florigen

    Nature

    (2011)
  • M. Alblova et al.

    Molecular basis of the 14-3-3 protein-dependent activation of yeast neutral trehalase Nth1

    Proc. Natl. Acad. Sci. U. S. A.

    (2017)
  • Cited by (0)

    View full text