Advertisement
Protein Structure and Folding| Volume 295, ISSUE 6, P1743-1753, February 07, 2020

Cryo-EM structure of human type-3 inositol triphosphate receptor reveals the presence of a self-binding peptide that acts as an antagonist

  • Caleigh M. Azumaya
    Footnotes
    Affiliations
    Department of Molecular Physiology and Biophysics, Vanderbilt University, School of Medicine, Nashville, Tennessee 37232
    Search for articles by this author
  • Emily A. Linton
    Footnotes
    Affiliations
    Department of Molecular Physiology and Biophysics, Vanderbilt University, School of Medicine, Nashville, Tennessee 37232
    Search for articles by this author
  • Caitlin J. Risener
    Affiliations
    Department of Molecular Physiology and Biophysics, Vanderbilt University, School of Medicine, Nashville, Tennessee 37232
    Search for articles by this author
  • Terunaga Nakagawa
    Affiliations
    Department of Molecular Physiology and Biophysics, Vanderbilt University, School of Medicine, Nashville, Tennessee 37232

    Center for Structural Biology, Vanderbilt University, School of Medicine, Nashville, Tennessee 37232

    Vanderbilt Brain Institute, Vanderbilt University, School of Medicine, Nashville, Tennessee 37232
    Search for articles by this author
  • Erkan Karakas
    Correspondence
    To whom correspondence should be addressed. Tel.: 615-343-4494;
    Affiliations
    Department of Molecular Physiology and Biophysics, Vanderbilt University, School of Medicine, Nashville, Tennessee 37232

    Center for Structural Biology, Vanderbilt University, School of Medicine, Nashville, Tennessee 37232
    Search for articles by this author
  • Author Footnotes
    1 Both authors contributed equally to this work.
Open AccessPublished:January 08, 2020DOI:https://doi.org/10.1074/jbc.RA119.011570
      Calcium-mediated signaling through inositol 1,4,5-triphosphate receptors (IP3Rs) is essential for the regulation of numerous physiological processes, including fertilization, muscle contraction, apoptosis, secretion, and synaptic plasticity. Deregulation of IP3Rs leads to pathological calcium signaling and is implicated in many common diseases, including cancer and neurodegenerative, autoimmune, and metabolic diseases. Revealing the mechanism of activation and inhibition of this ion channel will be critical to an improved understanding of the biological processes that are controlled by IP3Rs. Here, we report structural findings of the human type-3 IP3R (IP3R-3) obtained by cryo-EM (at an overall resolution of 3.8 Å), revealing an unanticipated regulatory mechanism where a loop distantly located in the primary sequence occupies the IP3-binding site and competitively inhibits IP3 binding. We propose that this inhibitory mechanism must differ qualitatively among IP3R subtypes because of their diverse loop sequences, potentially serving as a key molecular determinant of subtype-specific calcium signaling in IP3Rs. In summary, our structural characterization of human IP3R-3 provides critical insights into the mechanistic function of IP3Rs and into subtype-specific regulation of these important calcium-regulatory channels.

      Introduction

      Inositol 1,4,5-triphosphate receptors (IP3Rs)
      The abbreviations used are: IP3R
      inositol 1,4,5-triphosphate receptor
      ER
      endoplasmic reticulum
      IP3
      inositol 1,4,5-triphosphate
      PTEN
      phosphatase and tensin homolog
      CD
      cytoplasmic domain
      TMD
      transmembrane domain
      CTD
      C-terminal cytoplasmic domain
      CLD
      central linker domain
      β-TF
      β-trefoil domains
      ARM
      Armadillo solenoid domain
      LBD
      ligand-binding domain
      JD
      juxtamembrane domain
      NTD
      N-terminal domain
      MST
      microscale thermophoresis
      RyR-1
      rabbit type-1 ryanodine receptors
      β-ME
      β-mercaptoethanol
      GDN
      glyco-diosgenin
      LMN
      lauryl maltose neopentyl glycol
      TCEP
      tris(2-carboxyethyl)phosphine
      SEC
      size exclusion chromatography
      ITC
      isothermal titration calorimetry
      r.m.s. deviation
      root mean square deviation.
      are ligand-gated calcium (Ca2+) release channels localized predominantly in the endoplasmic reticulum (ER) membrane of all cell types (
      • Foskett J.K.
      • White C.
      • Cheung K.H.
      • Mak D.O.
      Inositol trisphosphate receptor Ca2+ release channels.
      ). Inositol 1,4,5-triphosphate (IP3) generated by phospholipase C upon G protein- or tyrosine kinase-coupled receptor activation binds to IP3Rs and opens the channel, leading to transfer of Ca2+ from the ER lumen to the cytoplasm (
      • Foskett J.K.
      • White C.
      • Cheung K.H.
      • Mak D.O.
      Inositol trisphosphate receptor Ca2+ release channels.
      ). Ca2+ released by IP3Rs act as universal messengers required to regulate diverse physiological processes including fertilization, muscle contraction, apoptosis, secretion, and synaptic plasticity (
      • Parys J.B.
      • De Smedt H.
      Inositol 1,4,5-trisphosphate and its receptors.
      ,
      • Berridge M.J.
      The inositol trisphosphate/calcium signaling pathway in health and disease.
      ). Deregulation of IP3Rs results in abnormal Ca2+ signaling, leading to a broad spectrum of pathologies including cancer, neurodegenerative, autoimmune, and metabolic diseases (
      • Mikoshiba K.
      Role of IP3 receptor signaling in cell functions and diseases.
      ).
      In mammals, there are three different subtypes (
      • Foskett J.K.
      • White C.
      • Cheung K.H.
      • Mak D.O.
      Inositol trisphosphate receptor Ca2+ release channels.
      ,
      • Parys J.B.
      • De Smedt H.
      Inositol 1,4,5-trisphosphate and its receptors.
      ,
      • Berridge M.J.
      The inositol trisphosphate/calcium signaling pathway in health and disease.
      ) of IP3Rs, which share 60–70% sequence identity, can form homo- or heterotetramers, exhibit different spatial expression profiles, and are involved in diverse signaling pathways. The type 3 receptors (IP3R-3s) are predominantly expressed in rapidly proliferating cells and are involved in taste perception and hair growth (
      • Sato-Miyaoka M.
      • Hisatsune C.
      • Ebisui E.
      • Ogawa N.
      • Takahashi-Iwanaga H.
      • Mikoshiba K.
      Regulation of hair shedding by the type 3 IP3 receptor.
      ,
      • Miura H.
      • Nakayama A.
      • Shindo Y.
      • Kusakabe Y.
      • Tomonari H.
      • Harada S.
      Expression of gustducin overlaps with that of type III IP3 receptor in taste buds of the rat soft palate.
      ,
      • Hisatsune C.
      • Yasumatsu K.
      • Takahashi-Iwanaga H.
      • Ogawa N.
      • Kuroda Y.
      • Yoshida R.
      • Ninomiya Y.
      • Mikoshiba K.
      Abnormal taste perception in mice lacking the type 3 inositol 1,4,5-trisphosphate receptor.
      ). Additionally, deregulation of IP3R-3 is implicated in diseases with deficiencies in cell fate decisions such as cancer and degenerative diseases (
      • Mendes C.C.
      • Gomes D.A.
      • Thompson M.
      • Souto N.C.
      • Goes T.S.
      • Goes A.M.
      • Rodrigues M.A.
      • Gomez M.V.
      • Nathanson M.H.
      • Leite M.F.
      The type III inositol 1,4,5-trisphosphate receptor preferentially transmits apoptotic Ca2+ signals into mitochondria.
      ,
      • Hayashi T.
      • Su T.P.
      Sigma-1 receptor chaperones at the ER-mitochondrion interface regulate Ca2+ signaling and cell survival.
      ,
      • Kuchay S.
      • Giorgi C.
      • Simoneschi D.
      • Pagan J.
      • Missiroli S.
      • Saraf A.
      • Florens L.
      • Washburn M.P.
      • Collazo-Lorduy A.
      • Castillo-Martin M.
      • Cordon-Cardo C.
      • Sebti S.M.
      • Pinton P.
      • Pagano M.
      PTEN counteracts FBXL2 to promote IP3R3- and Ca2+-mediated apoptosis limiting tumour growth.
      ). For example, the expression of the IP3R-3 is up-regulated in several cancer types including glioblastoma, breast, gastric, and colorectal cancer (
      • Mound A.
      • Vautrin-Glabik A.
      • Foulon A.
      • Botia B.
      • Hague F.
      • Parys J.B.
      • Ouadid-Ahidouch H.
      • Rodat-Despoix L.
      Downregulation of type 3 inositol (1,4,5)-trisphosphate receptor decreases breast cancer cell migration through an oscillatory Ca2+ signal.
      ,
      • Kang S.S.
      • Han K.S.
      • Ku B.M.
      • Lee Y.K.
      • Hong J.
      • Shin H.Y.
      • Almonte A.G.
      • Woo D.H.
      • Brat D.J.
      • Hwang E.M.
      • Yoo S.H.
      • Chung C.K.
      • Park S.H.
      • Paek S.H.
      • Roh E.J.
      • et al.
      Caffeine-mediated inhibition of calcium release channel inositol 1,4,5-trisphosphate receptor subtype 3 blocks glioblastoma invasion and extends survival.
      ,
      • Sakakura C.
      • Hagiwara A.
      • Fukuda K.
      • Shimomura K.
      • Takagi T.
      • Kin S.
      • Nakase Y.
      • Fujiyama J.
      • Mikoshiba K.
      • Okazaki Y.
      • Yamagishi H.
      Possible involvement of inositol 1,4,5-trisphosphate receptor type 3 (IP3R3) in the peritoneal dissemination of gastric cancers.
      ,
      • Shibao K.
      • Fiedler M.J.
      • Nagata J.
      • Minagawa N.
      • Hirata K.
      • Nakayama Y.
      • Iwakiri Y.
      • Nathanson M.H.
      • Yamaguchi K.
      The type III inositol 1,4,5-trisphosphate receptor is associated with aggressiveness of colorectal carcinoma.
      ). Furthermore, many tumor suppressors and oncoproteins such as protein kinase B, protein phosphatase 2A, promyelocytic leukemia protein, phosphatase and tensin homolog (PTEN), and BRCA1-associated protein 1 tightly regulate the stability and activity of IP3R-3s (
      • Kuchay S.
      • Giorgi C.
      • Simoneschi D.
      • Pagan J.
      • Missiroli S.
      • Saraf A.
      • Florens L.
      • Washburn M.P.
      • Collazo-Lorduy A.
      • Castillo-Martin M.
      • Cordon-Cardo C.
      • Sebti S.M.
      • Pinton P.
      • Pagano M.
      PTEN counteracts FBXL2 to promote IP3R3- and Ca2+-mediated apoptosis limiting tumour growth.
      ,
      • Bononi A.
      • Giorgi C.
      • Patergnani S.
      • Larson D.
      • Verbruggen K.
      • Tanji M.
      • Pellegrini L.
      • Signorato V.
      • Olivetto F.
      • Pastorino S.
      • Nasu M.
      • Napolitano A.
      • Gaudino G.
      • Morris P.
      • Sakamoto G.
      • et al.
      BAP1 regulates IP3R3-mediated Ca2+ flux to mitochondria suppressing cell transformation.
      ,
      • Giorgi C.
      • Ito K.
      • Lin H.K.
      • Santangelo C.
      • Wieckowski M.R.
      • Lebiedzinska M.
      • Bononi A.
      • Bonora M.
      • Duszynski J.
      • Bernardi R.
      • Rizzuto R.
      • Tacchetti C.
      • Pinton P.
      • Pandolfi P.P.
      PML regulates apoptosis at endoplasmic reticulum by modulating calcium release.
      ,
      • Szado T.
      • Vanderheyden V.
      • Parys J.B.
      • De Smedt H.
      • Rietdorf K.
      • Kotelevets L.
      • Chastre E.
      • Khan F.
      • Landegren U.
      • Söderberg O.
      • Bootman M.D.
      • Roderick H.L.
      Phosphorylation of inositol 1,4,5-trisphosphate receptors by protein kinase B/Akt inhibits Ca2+ release and apoptosis.
      ). Inhibiting IP3R-3 degradation in PTEN-regulated cancers was shown to be a valid therapeutic strategy (
      • Bononi A.
      • Giorgi C.
      • Patergnani S.
      • Larson D.
      • Verbruggen K.
      • Tanji M.
      • Pellegrini L.
      • Signorato V.
      • Olivetto F.
      • Pastorino S.
      • Nasu M.
      • Napolitano A.
      • Gaudino G.
      • Morris P.
      • Sakamoto G.
      • et al.
      BAP1 regulates IP3R3-mediated Ca2+ flux to mitochondria suppressing cell transformation.
      ). Although IP3R-3s are responsible for regulating distinct biological processes compared with types 1 and 2, it is unclear whether there is any mechanistic difference in their operation.
      IP3Rs function as signaling hubs where signals from different pathways and metabolic states are integrated to allosterically modulate IP3R gating. Receptor activity is tightly controlled by many factors including second messengers (e.g. IP3, Ca2+), other small molecules (e.g. ATP), modulatory proteins, and posttranslational modifications such as phosphorylation and ubiquitination. Despite recent advances in the structural studies of IP3Rs, molecular understanding of receptor gating and regulation remains largely unknown.
      Structural investigation of IP3Rs was pioneered by using the IP3R-1 obtained from native tissues, due to its abundant expression in the cerebellum and well-established purification strategy (
      • Fan G.
      • Baker M.L.
      • Wang Z.
      • Baker M.R.
      • Sinyagovskiy P.A.
      • Chiu W.
      • Ludtke S.J.
      • Serysheva I.I.
      Gating machinery of InsP3R channels revealed by electron cryomicroscopy.
      ,
      • da Fonseca P.C.
      • Morris S.A.
      • Nerou E.P.
      • Taylor C.W.
      • Morris E.P.
      Domain organization of the type 1 inositol 1,4,5-trisphosphate receptor as revealed by single-particle analysis.
      ,
      • Hamada K.
      • Terauchi A.
      • Mikoshiba K.
      Three-dimensional rearrangements within inositol 1,4,5-trisphosphate receptor by calcium.
      ,
      • Jiang Q.X.
      • Thrower E.C.
      • Chester D.W.
      • Ehrlich B.E.
      • Sigworth F.J.
      Three-dimensional structure of the type 1 inositol 1,4,5-trisphosphate receptor at 24 Å resolution.
      ,
      • Sato C.
      • Hamada K.
      • Ogura T.
      • Miyazawa A.
      • Iwasaki K.
      • Hiroaki Y.
      • Tani K.
      • Terauchi A.
      • Fujiyoshi Y.
      • Mikoshiba K.
      Inositol 1,4,5-trisphosphate receptor contains multiple cavities and L-shaped ligand-binding domains.
      ,
      • Serysheva I.I.
      • Bare D.J.
      • Ludtke S.J.
      • Kettlun C.S.
      • Chiu W.
      • Mignery G.A.
      Structure of the type 1 inositol 1,4,5-trisphosphate receptor revealed by electron cryomicroscopy.
      ,
      • Nakade S.
      • Rhee S.K.
      • Hamanaka H.
      • Mikoshiba K.
      Cyclic AMP-dependent phosphorylation of an immunoaffinity-purified homotetrameric inositol 1,4,5-trisphosphate receptor (type I) increases Ca2+ flux in reconstituted lipid vesicles.
      ,
      • Supattapone S.
      • Worley P.F.
      • Baraban J.M.
      • Snyder S.H.
      Solubilization, purification, and characterization of an inositol trisphosphate receptor.
      ,
      • Fan G.
      • Baker M.R.
      • Wang Z.
      • Seryshev A.B.
      • Ludtke S.J.
      • Baker M.L.
      • Serysheva I.I.
      Cryo-EM reveals ligand induced allostery underlying InsP3R channel gating.
      ). Recently, cryo-EM structures of recombinant IP3R-3 have been reported and the global architectures of both IP3R-1 and IP3R-3 share common features (
      • Paknejad N.
      • Hite R.K.
      Structural basis for the regulation of inositol trisphosphate receptors by Ca(2+) and IP3.
      ). Due to the large size of the IP3Rs, different structures contain nonoverlapping information resulting, primarily, from variations of local resolution within the 3D map. In fact, the atomic view of the complex is still incomplete. Published structures in the unliganded (apo) and liganded states provide a basis to develop mechanistic hypotheses on the channel gating. However, the current conformational ensemble falls short in revealing the full gating cycle, modulation, as well as mechanism of inhibition by known chemical reagents.
      Here we present a cryo-EM structure of the human IP3R-3 (hIP3R-3) in a ligand-free conformation revealing a loop extending from the regulatory ARM2 domain that occupies the IP3-binding site and thus may function as a regulator of IP3 binding. In addition, our structure identified previously unresolved local structures of the complex, and the location of lipid-binding sites in the transmembrane domain. Collectively, our structural characterization of the hIP3R-3 provides novel insight into the mechanistic function of IP3Rs.

      Results

      Structure of hIP3R-3

      We expressed recombinant hIP3R-3 using the Sf9 insect cell/baculovirus expression system and purified detergent-solubilized protein in the absence of any known ligands (Fig. S1). After initial analysis of the protein sample using negative-stain EM, we solved its structure using cryo-EM to an overall resolution of 3.8 Å (Fig. 1 and Figs. S2–S4). We used high resolution crystal structures for the ligand-binding core (PDB IDs 3JRR and 3UJ4) to interpret the map and modeled the rest of the structure manually (
      • Chan J.
      • Yamazaki H.
      • Ishiyama N.
      • Seo M.D.
      • Mal T.K.
      • Michikawa T.
      • Mikoshiba K.
      • Ikura M.
      Structural studies of inositol 1,4,5-trisphosphate receptor: coupling ligand binding to channel gating.
      ,
      • Seo M.D.
      • Velamakanni S.
      • Ishiyama N.
      • Stathopulos P.B.
      • Rossi A.M.
      • Khan S.A.
      • Dale P.
      • Li C.
      • Ames J.B.
      • Ikura M.
      • Taylor C.W.
      Structural and functional conservation of key domains in InsP3 and ryanodine receptors.
      ). The overall structure of the hIP3R-3 expressed in Sf9 cells is consistent with the structures of hIP3R-3 expressed in HEK GnTI(−) cells, and very similar to the structure of rat IP3R-1 purified from native tissues (
      • Fan G.
      • Baker M.L.
      • Wang Z.
      • Baker M.R.
      • Sinyagovskiy P.A.
      • Chiu W.
      • Ludtke S.J.
      • Serysheva I.I.
      Gating machinery of InsP3R channels revealed by electron cryomicroscopy.
      ,
      • Fan G.
      • Baker M.R.
      • Wang Z.
      • Seryshev A.B.
      • Ludtke S.J.
      • Baker M.L.
      • Serysheva I.I.
      Cryo-EM reveals ligand induced allostery underlying InsP3R channel gating.
      ,
      • Paknejad N.
      • Hite R.K.
      Structural basis for the regulation of inositol trisphosphate receptors by Ca(2+) and IP3.
      ).
      Figure thumbnail gr1
      Figure 1Cryo-EM structure of hIP3R-3. A and B, density map of the hIP3R-3 viewed along the membrane plane (A) and from cytosol (B). Each domain in one of the subunits is colored differently. The other subunits are colored in different shades of gray. Additional density occupying the IP3 binding. C, ribbon representation of the hIP3R-3 subunit highlighting the domain architecture. Zinc ion is shown as gray sphere. D, density map around the IP3-binding site is shown in gray (transparent) along with the additional density colored in magenta. The residues forming the pocket are shown in sticks. E, domain boundaries of hIP3R-3 subunits. Domains are colored as in panels A–C.
      Subunits that form the tetrameric ion channel can be divided into 3 regions: the large, N-terminal cytoplasmic domain (CD), the channel-forming transmembrane domain (TMD), and the C-terminal cytoplasmic domain (CTD) (Fig. 1C). The CD of each subunit resembles a tripod with a hinge-like central linker domain (CLD) (residues 790–1100 and 1587–1697) connected to 3 Armadillo solenoid domains (ARM1–3) (Fig. 1C). The CLD is located at the outer perimeter of the tetrameric receptor. The N-terminal domain (ARM1) extends toward the central 4-fold symmetry axis and connects to 2 contiguous β-trefoil domains (β-TF1 and β-TF2), forming the ligand-binding domain (LBD). β-TF1 of one subunit interacts with β-TF2 of the neighboring subunit forming a rim around the 4-fold symmetry axis. The second ARM domain (ARM2) bulges from the CLD and is oriented parallel to the membrane surface. It interacts with ARM1 of the neighboring subunit and forms the outer periphery of the receptor together with the CLD. The third ARM domain (ARM3) connects the cytoplasmic domains to a juxtamembrane domain (JD) positioned at the cytoplasmic face of the TMD. The JD is formed by assembly of two fragments separated by the TMD. A U-motif composed of a β-hairpin and a helix-turn-helix motif located at the C-terminal end of the ARM3 domain encapsulates a latch-like domain extending from the C-terminal end of the TMD. The JD is further stabilized by a H2C2 zinc finger domain formed by the residues Cys-2538, Cys-2541, His-2558, His-2563, and a zinc ion (Fig. 1C).

      IP3-binding site is occupied by a loop extending from ARM2

      We observed substantive density at the IP3-binding site, despite our initial intention to obtain the structure in a ligand-free conformation (Fig. 1, A–D). The density for the residues forming the binding pocket was well-resolved indicating that the additional density is from a potential ligand that occupies the IP3-binding site; hereafter we will refer to this density as “ligand-like” density (Fig. 1D). To improve the quality of the density at the IP3-binding site, we first treated each individual subunit as a single particle and artificially expanded the dataset by symmetry expansion around the C4 symmetry axis that increased the number of particles 4-fold from 82,511 to 330,044 (Fig. S3) (
      • Scheres S.H.W.
      Processing of Structurally Heterogeneous Cryo-EM Data in RELION.
      ). Then, we performed partial signal subtraction from the experimental images to reduce the signal to a region that would only cover the first ∼1,800 residues (IP3R-3 NTD) including the LBD as well as ARM1, ARM2, CLD, and part of ARM3 domains of a single protomer (Fig. S3) (
      • Scheres S.H.W.
      Processing of Structurally Heterogeneous Cryo-EM Data in RELION.
      ). The subtracted particles were then classified into six 3D classes using a mask that covers the IP3R-3 NTD without performing angular or translational alignment (Fig. S3). Although all six maps from this classification scheme contained the ligand-like density at the IP3-binding site, two of the classes revealed continuous density connecting the “ligand-like” to the first and second helices of ARM2 (αARM2–1 and αARM2–2) (Fig. 2A and Fig. S5, A and B). We observed similar results from the 3D classifications performed using different strategies as discussed under “Experimental procedures.” Further investigation of the surrounding area in the model led us to surmise this extended, connecting ligand-like density could, in fact, be an unmodeled loop of the ARM2 domain. In our model, based on the 3D reconstruction of the intact receptor, residues connecting αARM2–1 to αARM2–2 are not modeled due to lack of interpretable density potentially resulting from the intrinsic flexibility of this loop (Fig. 2, B and C). This loop was also left unmodeled, presumably due to flexibility, from previously published IP3R-3 structures (
      • Paknejad N.
      • Hite R.K.
      Structural basis for the regulation of inositol trisphosphate receptors by Ca(2+) and IP3.
      ). Within this region there are two patches enriched in acidic residues (Fig. 2E). Therefore, it is plausible that this loop forms a self-binding peptide (SBP) extending toward the IP3-binding site and bringing one or both of these patches in close proximity to basic residues in the pocket of the IP3-binding site (Fig. 2B–D). SBPs are peptide segments that specifically recognize and interact with their cognate targets, while being incorporated to the target in the primary sequence via a flexible polypeptide linker (
      • Yang C.
      • Zhang S.
      • He P.
      • Wang C.
      • Huang J.
      • Zhou P.
      Self-binding peptides: folding or binding?.
      ).
      Figure thumbnail gr2
      Figure 2IP3-binding site is occupied by a loop extending from ARM2. A, density maps of hIP3R-3 NTD class 5 (left) and class 3 (right) after focused 3D classification. Arrows point to the density extending from ARM2 domain. B, ribbon representation of hIP3R-3 NTD along with transparent surface rendering of the density map of hIP3R-3 NTD- class 5. Dashed lines represent the putative path for the Cα atoms of the residues forming the loop between αARM2–1 and αARM2–2. C, close up view of the IP3-binding site of the SBP-bound hIP3R-3 along with a docking pose of IP3, placed based on the alignment of the LBDs of IP3- and SBP-bound hIP3R-3s. D, electrostatic surface representation of the IP3-binding site with a modeled loop in red. E, sequence alignment of hIP3R subtypes around the sequence covering αARM2–1 and αARM2–2. Cylinders represent α-helices. Dashed lines indicate the disordered region in the reconstruction of entire receptor with C4 symmetry. Acidic residues within the loop are colored red.
      Although the presence of a density representing a loop extending from the ARM2 domain is unambiguous, modeling of specific amino acids forming the loop was not possible due to weak features observed resulting from the flexibility of the loop and ARM2. Each 3D class had a different ARM2 arrangement relative to the rest of the CD through a rigid body rotation on a pivotal point where ARM2 is connected to the CLD (Fig. S5C). In addition, density at the IP3-binding site is not uniform among different 3D classes implying a dynamic interaction between the SBP and the IP3-binding site (Fig. S5, A and B).

      IP3R-3 SBP competes against IP3 binding

      We reasoned the SBP loop would compete against IP3 binding if it is interacting with the IP3-binding site. To test this hypothesis, we took two approaches. In the first approach, we expressed and purified the hIP3R LBD (residues 4–602) and a larger N-terminal domain (NTD; residues 4–1799), containing ARM2, as GFP fusion proteins and performed microscale thermophoresis (MST) experiments to measure the binding affinity of IP3 (Fig. 3, A and B). IP3 affinity for the NTD was over 7-fold lower than for the LBD alone (Kd = 1.31 ± 0.46 and 0.18 ± 0.028 μm, respectively) (Fig. 3B). Deletion of the putative SBP (residues 1133–1155) from the NTD construct increased the affinity for IP3 (Kd = 0.25 ± 0.09 μm), not significantly different from that of the LBD alone (Fig. 3B). Furthermore, mutagenesis of the four glutamate residues at both acidic patches to alanine (E1136A, E1137A, E1153A, and E1154A) in the putative SBP of the full-length NTD protein caused a similar increase in the IP3 affinity (Kd = 0.44 ± 0.20 μm). Taken together, these observations support the hypothesis that residues 1133–1155 of hIP3R-3 form a SBP that competes for binding of the major agonist, soluble IP3.
      Figure thumbnail gr3
      Figure 3The SBP competes against IP3 binding. A, structural and schematic representation of the proteins used for binding assays. Changes in the SBP sequence was shown using dashes for deletion construct and in red for glutamate to alanine mutations. B, MST analysis of IP3 interaction with hIP3R-3-LBD (green), hIP3R-3-NTD (blue), hIP3R-3-NTD-ΔSBP (red), and hIP3R-3-NTD-E1136–1137-1153–1154A (cyan). Error bars represent standard deviations from three individual repeat measurements. C, calorimetric titration of IP3 into hIP3R-3 LBD, hIP3R-3 LBD + SBP, and hIP3R-3 LBD + SBP-E1136–1137-1153–1154A (upper panels) and integrated heat as a function of IP3/protein ratio (lower panels). Calculated Kd values are shown for each panel.
      In the second approach to test this hypothesis, we prepared a construct where the SBP was fused via a flexible linker to the C-terminal end of the LBD so it would be in close spatial proximity to the IP3-binding site (Fig. 3A). We hypothesized this would behave like a gain-of-function construct, where the LBD-SBP fusion would have lower affinity for IP3, similar to the NTD construct used above. We performed ITC experiments to test if the presence of the SBP affects the protein's affinity for IP3. Similar to the MST experiments, the LBD's affinity for IP3 decreased nearly 3-fold in the presence of the SBP (Kd = 0.19 ± 0.02 to 0.52 ± 0.02 μm) (Fig. 3C). This effect was again abolished when the acidic residues in the SBP were mutated to alanines (Fig. 3, A–C). These results further support the hypotheses that the SBP competes against IP3 binding and that the acidic residues are important for this effect, as suggested by the models of our IP3R-3 structure.

      IP3R-3 SBP binding

      When compared with the apo-LBD (PDB ID 6DQJ) (
      • Paknejad N.
      • Hite R.K.
      Structural basis for the regulation of inositol trisphosphate receptors by Ca(2+) and IP3.
      ), with no visible density at the IP3-binding site, the LBD of the SBP-bound hIP3R-3 adopts a very similar overall conformation with a few local differences at the loops forming the IP3-binding site at the β-TF2 domain (Fig. 4A). The most apparent difference involves the loop formed by Leu-269 and Arg-270, which is positioned closer to the SBP density, and the side chain of Leu-269, which points in the opposite direction (Fig. 4A). This slight movement does not seem to affect the overall arrangement of the IP3R-3 CD, as the whole CD can be superimposed with an r.m.s.d value of 1.1 Å over 1,728 aligned residues. However, there is a subtle but noticeable counter-clockwise rotation (∼3°) of the entire tetrameric CD relative to the TMD in the SBP-bound hIP3R-3 compared with the apo-hIP3R-3, when the TMDs of both structures are aligned (Fig. 4C).
      Figure thumbnail gr4
      Figure 4Structural comparison of SBP-bound IP3R-3 with apo and IP3-bound hIP3Rs. A and B, β-TF1 and β-TF2 domains of the SBP-bound hIP3R LBD are superposed onto β-TF1 and β-TF2 domains of (A) apo hIP3R-3 (PDB ID 6DQJ, shown in blue) and (B) IP3-bound -hIP3R-3 (PDB ID 6DQN, shown in orange). β-TF2 and ARM1 domains of the SBP-bound hIP3R LBD are colored in yellow and cyan, respectively. β-TF1 domains are not shown in the figure for clarity. Residues forming the IP3-binding pocket are shown in stick representation. IP3 is colored black and magenta mesh represents the SBP density. C and D, comparison of the CDs of SBP-bound hIP3R-3 with (C) apo-hIP3R-3 (blue) and (D) IP3-bound -hIP3R-3 (orange) after superposing the TMDs. Subunits of the SBP-bound hIP3R-3 are colored in cyan, salmon, green, and yellow.
      Previous structures of hIP3R-3 in complex with IP3 revealed two different LBD conformations; apo-like class 1 and class 2 with substantial conformational changes (
      • Paknejad N.
      • Hite R.K.
      Structural basis for the regulation of inositol trisphosphate receptors by Ca(2+) and IP3.
      ). The LBD in SBP-bound conformation is more similar to the apo-like class-1 structure (PDB ID 6DQN, Fig. 4B). Orientation of the loop formed by Leu-269 and Arg-270 is highly similar in both structures suggesting that conformational changes of this loop are coupled to ligand binding (Fig. 4B). The arrangement of the tetrameric CD relative to the TMD is also comparable in both structures (Fig. 4D). In contrast, when compared with the ARM1 domain in the SBP-bound structure, the class 2 IP3-bound structure (PDB ID 6DQV) ARM1 domain is rotated by ∼20° toward to the β-TF2 domain (Fig. S6). This rotation leads to overlap of residues in the ARM1 domain of the class 2 structure with our SBP density. Therefore, it is not possible for the LBD to adopt a similar conformation when the SBP occupies the IP3-binding site due to steric hindrance, suggesting that the SBP would act as an antagonist.

      Transmembrane domain

      The TMD structure is consistent with previous IP3R structures and has the overall architecture of voltage-gated ion channels with a central pore domain surrounded by voltage sensor-like domains at the periphery (Figs. 1C and 5). Unlike voltage-gated ion channels, we observe two additional helices (S1′ and S1″) per subunit penetrating through the membrane from the luminal side, similar to those observed in the previous cryo-EM studies of hIP3R-3s as well as rabbit type-1 ryanodine receptors (RyR-1s) (
      • Paknejad N.
      • Hite R.K.
      Structural basis for the regulation of inositol trisphosphate receptors by Ca(2+) and IP3.
      ,
      • des Georges A.
      • Clarke O.B.
      • Zalk R.
      • Yuan Q.
      • Condon K.J.
      • Grassucci R.A.
      • Hendrickson W.A.
      • Marks A.R.
      • Frank J.
      Structural basis for gating and activation of RyR1.
      ). Thus, these auxiliary TM helices seem to be a common feature of intracellular calcium release channels. Primary sequences of these two helices are the most diverse region within the TMD among 3 subtypes of IP3Rs and could potentially be involved in subtype-specific regulation and/or localization of the IP3Rs.
      Figure thumbnail gr5
      Figure 5Structure of the TMD. A and C, ribbon diagram of the hIP3R-3 TMD viewed from the cytoplasm (A) and ER lumen (C) with nonprotein density countered at 4 σ representing bound-lipid molecules. B and D, the binding site for L-1 (B) and L-2 (D) viewed in the same orientation as in panels A and C, respectively. Density for the protein and the lipid molecules are shown in gray and red, respectively. Residues that are involved in L-1 binding are shown as sticks. A putative phosphatidylethanolamine molecule in stick representation was placed into the L-1 density for illustrative purposes, although it was not built in the actual structural model.
      Through the ion permeation path of the channel, from the cytoplasmic side, there is an upper vestibule, the narrowest constriction of the channel, and a lower vestibule followed by an architecture similar to the selectivity filter seen in potassium channels (Fig. S7). In agreement with a closed channel conformation in the resting state, the shortest pore diameter along the channel was 1.1 Å where residues Phe-2513 and Ile-2522 are located (
      • Smart O.S.
      • Neduvelil J.G.
      • Wang X.
      • Wallace B.A.
      • Sansom M.S.
      HOLE: a program for the analysis of the pore dimensions of ion channel structural models.
      ,
      • Marcus Y.
      Ionic radii in aqueous solutions.
      ; Fig. S7). At the lower vestibule side of this constriction, there is a π-helix (residues 2501–2509) located at the middle of the S6 helix (Fig. S7). As suggested previously, transition from π- to α-helix within this region is potentially coupled to gating similar to TRPV6 channels where channel opening is accompanied by a local α- to π-helical transition in S6 (
      • Paknejad N.
      • Hite R.K.
      Structural basis for the regulation of inositol trisphosphate receptors by Ca(2+) and IP3.
      ,
      • McGoldrick L.L.
      • Singh A.K.
      • Saotome K.
      • Yelshanskaya M.V.
      • Twomey E.C.
      • Grassucci R.A.
      • Sobolevsky A.I.
      Opening of the human epithelial calcium channel TRPV6.
      ).
      We observed two strong, nonprotein densities per subunit at the TMD. The true identity of the molecules occupying these positions cannot be determined with certainty from the current data, but they potentially derive from either nonannular lipid molecules co-purified with the receptor or well-ordered detergent molecules (Fig. 5). In either case, these sites are likely to be occupied by lipids in biological membranes, as they are embedded in the TMD core and form extensive interface with the protein residues. The density located in the cytoplasmic leaflet of the bilayer is located at the cavity formed by the S3, S4, and S4–5 helices and is in the vicinity of residues Tyr-2322, Ile-2349, Tyr-2350, and Phe-2356 (Fig. 5, A and B). The tip of the density facing the cytoplasmic side extends toward Glu-2353, suggesting that the lipid molecule contains a positively charged head group that can form a salt bridge with the side chain of Glu-2353. The second lipid density is located at the interface of three subunits in the luminal leaflet of the bilayer and could be critical for proper assembly (Fig. 5, C and D). The binding site is formed by S1 and S1′ of one subunit together with the P helix and S6 helix of two neighboring subunits (Fig. 5D).

      C-terminal cytoplasmic domain

      Unlike RyRs, the C-terminal ends of IP3Rs extend through the central 4-fold axis and form a left-handed helical bundle at the core of the receptor (
      • Fan G.
      • Baker M.L.
      • Wang Z.
      • Baker M.R.
      • Sinyagovskiy P.A.
      • Chiu W.
      • Ludtke S.J.
      • Serysheva I.I.
      Gating machinery of InsP3R channels revealed by electron cryomicroscopy.
      ). In our structure, density for the CTD is less resolved compared with the rest of the receptor, but in sufficient quality to model a polyalanine peptide that forms a left-handed coiled-coil motif (Fig. 6, A and B). Connection of the coiled-coiled motif to the JD is only visible at very low threshold levels, and was not built into the model. The C-terminal side of the helices forming the coiled-coiled domain extend toward the β-TF ring and interact with the β-TF2 domain of the neighboring subunit (Fig. 6). The interaction is mediated through a hydrophobic patch formed by Val-287, Val-288, Leu-303, Ile-363, Leu-366, and Leu-392 on the surface of the β-TF2 domain and potentially involves residues Leu-2660, Gly-2661, Phe-2662, Val-2663, Asp-2664, and Val-2665 at the C-terminal end of the receptor (Fig. 6, C and D).
      Figure thumbnail gr6
      Figure 6Coupling between the N- and C-terminal domains of hIP3R-3. A, density map of hIP3R-3 viewed from the cytoplasmic side. Density for each subunit is colored differently. B, close up view of the map in panel A focusing on the C-terminal coiled-coil domain. C, surface representation of the of the β-TF1 and β-TF2 domains colored using the YRB coloring scheme (
      • Hagemans D.
      • van Belzen I.A.
      • Morán Luengo T.
      • Rudiger S.G.
      A script to highlight hydrophobicity and charge on protein surfaces.
      ) (yellow, carbon atoms not connected to nitrogen and oxygen atoms; red, negatively charged atoms; blue, positively charged atoms; and gray, remaining atoms) along with the cryo-EM map (gray mesh) of the C-terminal end of the neighboring subunit to emphasize the hydrophobic nature of the interaction. The residues of the β-TF2 domain in close distance to the C-terminal end of the neighboring subunit are shown in sticks and labeled. D, sequence alignment of hIP3R subtypes around the C-terminal end of the proteins. Underlined sequences represent the predicted coiled-coil motifs. Residues shown in blue form the hydrophobic patch proposed to interact with the hydrophobic surface of the β-TF2 domain of the neighboring subunit.

      Discussion

      The identified direct interaction between the SBP and the ligand-binding site may have a critical physiological role in IP3R-3 activity and regulation. By occupying the IP3-binding site, the SBP reduces the sensitivity of the IP3R-3 to its physiological agonist IP3. Similar mechanisms of receptor regulation were observed for other protein families as well. For example, fibroblast growth factor receptor autoinhibition is mediated by electrostatic interaction of a subregion rich in acidic residues, known as an acid box, with the heparin-binding site of the same subunit, reducing receptor affinity for heparin and fibroblast growth factor (
      • Kalinina J.
      • Dutta K.
      • Ilghari D.
      • Beenken A.
      • Goetz R.
      • Eliseenkova A.V.
      • Cowburn D.
      • Mohammadi M.
      The alternatively spliced acid box region plays a key role in FGF receptor autoinhibition.
      ). Of note, heparin is an antagonist of IP3Rs and interacts with the IP3-binding site.
      Among the IP3R subtypes, IP3R-3 has the lowest affinity to IP3. The subtype differences in IP3 affinity are mainly attributed to sequence variations at the β-TF1 domain (also known as the suppressor domain) (
      • Iwai M.
      • Michikawa T.
      • Bosanac I.
      • Ikura M.
      • Mikoshiba K.
      Molecular basis of the isoform-specific ligand-binding affinity of inositol 1,4,5-trisphosphate receptors.
      ). However, the current data indicate that the SBP also has a role in modulating binding affinity because of sequence variability within the SBP among the IP3R subtypes, especially in the number and positioning of the acidic residues (Fig. 2E). In addition, we speculate that the SBP is a plausible target for many proteins that are known to modulate the sensitivity to IP3. For example, phosphorylation of the Ser/Thr residues within the SBP would increase the net negative charge on the SBP and potentially lead to tighter binding to the IP3-binding site reducing the receptor sensitivity to IP3. Moreover, any protein that interacts with the SBP may restrict its interaction with the IP3-binding site, sensitizing the IP3R to IP3.
      In addition to a potential physiological role in regulation of IP3R activity, identification of the SBP as a competitive inhibitor of IP3 binding opens potential avenues in development of pharmacological agents targeting these important families of proteins. Further research to identify the structural determinants of the SBP interaction with the IP3-binding site will pave the road for development of novel inhibitors of IP3Rs.
      One puzzling question is why the additional density from the SBP at the IP3-binding site was not observed in a previously reported ligand-free hIP3R-3 structure. One plausible explanation is that the larger number of particles used in our study (82,511 compared with 26,325) provided additional information that permitted the resolution of the SBP. In addition, this could arise from technical differences, such as the protein expression system (mammalian versus insect cells), purification methods, and/or sample preparation for cryo-EM analysis. Furthermore, cryo-EM studies of large macromolecular complexes provide opportunities to identify novel features through different image-processing strategies even with samples prepared in similar conditions. IP3Rs are regulated by numerous factors, some of which are not well-understood at the moment, and exhibit multiple conformational rearrangements. Thus, it is likely that further structural studies of IP3Rs will continue to uncover additional features providing further functional insights.
      In conclusion, the data presented here reveal a previously unanticipated regulatory mechanism of IP3R where a loop distantly located from the LBD in the primary sequence occupies the IP3-binding site and competitively inhibits IP3 binding. Regulation by the SBP is likely to confer subtype-specific biological function to IP3-mediated calcium signaling due to divergence in the loop sequence among members of the IP3R family. Our structural data will facilitate design of modifications on the SBP of the intact receptors to functionally test their effect in channel activity to determine the molecular mechanism of the SBP regulation and its physiological role.

      Experimental procedures

      Expression and purification of hIP3R-3

      The gene encoding hIP3R-3 (accession number BC172406) was purchased from Dharmacon (
      • Strausberg R.L.
      • Feingold E.A.
      • Grouse L.H.
      • Derge J.G.
      • Klausner R.D.
      • Collins F.S.
      • Wagner L.
      • Shenmen C.M.
      • Schuler G.D.
      • Altschul S.F.
      • Zeeberg B.
      • Buetow K.H.
      • Schaefer C.F.
      • Bhat N.K.
      • Hopkins R.F.
      • et al.
      Generation and initial analysis of more than 15,000 full-length human and mouse cDNA sequences.
      ), subcloned (residues 4–2671) with C-terminal OneStrep tag into pFL vector, and incorporated into baculovirus using the Multibac expression system (
      • Fitzgerald D.J.
      • Berger P.
      • Schaffitzel C.
      • Yamada K.
      • Richmond T.J.
      • Berger I.
      Protein complex expression by using multigene baculoviral vectors.
      ). Sf9 cells (4 × 106 cells/ml) infected with the baculovirus were harvested by centrifugation (4,000 × g) 48 h after infection. Cells resuspended in a lysis buffer of 200 mm NaCl, 40 mm Tris-HCl, pH 8.0, 2 mm EDTA, pH 8.0, 10 mm β-mercaptoethanol (β-ME), 1 mm phenylmethylsulfonyl fluoride were lysed using Avastin EmulsiFlex-C3. The cell lysate was centrifuged at 6,000 × g for 20 min and the membrane was pelleted by centrifugation at 40,000 rpm (Ti45 rotor) for 1 h. Membrane pellets were resuspended and homogenized in ice-cold resuspension buffer (200 mm NaCl, 40 mm Tris-HCl, pH 8.0, 2 mm EDTA, pH 8.0, 10 mm β-ME), and solubilized using 0.5% lauryl maltose neopentyl glycol (LMNG) and 0.1% glyco-diosgenin (GDN) at a membrane concentration of 100 mg/ml. After 4 h of stirring, the insoluble material was separated by centrifugation at 40,000 rpm (Ti45 rotor) for 1 h and the supernatant was passed through Strep-XT-Superflow resin (IBA Biotagnology). The resin was washed with the wash buffer composed of 200 mm NaCl, 20 mm Tris-HCl, pH 8.0, 1 mm EDTA, 10 mm β-ME, 0.005% LMN, 0.005% GDN, and the protein was eluted using the wash buffer supplemented with 100 mmd-biotin, pH 8.2. Protein was further purified by size exclusion chromatography using Superose 6 (10/300 GL, GE Healthcare) equilibrated with 200 mm NaCl, 20 mm Tris-HCl, pH 8.0, 1 mm EDTA, pH 8.0, 2 mm TCEP, 0.005% LMN, and 0.005% GDN. Fractions corresponding to hIP3R-3 were concentrated to 2.3 mg/ml, centrifuged at 70,000 rpm using a S110-AT rotor (Thermo Scientific) for 10 min, and used immediately for cryo-EM imaging. After centrifugation, concentration of the protein decreased to 1.3 mg/ml.

      Negative stain data collection and analysis for hIP3R-3

      Four hundred mesh copper grids were coated with carbon. 4 μl of 0.05 mg/ml of hIP3R-3 was applied to each glow discharged grid (2 min on Quorum K1000X) and allowed to absorb for 30 s. The grid was blotted on filter paper, washed twice in MilliQ water, and negatively stained with 0.75% (w/v) uranyl formate. Images were recorded on a 4k × 4k CCD camera (Gatan) using an FEI TF20 transmission electron microscope operated at 200 keV. All images were collected at 50,000 magnification in low dose mode using SerialEM automated collection mode at a defocus of −1.5 μm (
      • Mastronarde D.N.
      Automated electron microscope tomography using robust prediction of specimen movements.
      ). Image processing was performed using the RELION software package (
      • Scheres S.H.
      RELION: implementation of a Bayesian approach to cryo-EM structure determination.
      ). Particles were automatically picked using templates generated from 822 manually picked particles. 100 2D class averages were generated from 12,227 particles using 25 iterations of 2D classification and alignment in RELION (Fig. S2A).

      Cryo-EM sample preparation and data collection for hIP3R-3

      2.0 μl of 1.3 mg/ml of hIP3R-3 was applied to a 200-mesh C-flat holey carbon 2/1 grid (Protochips) that was glow discharged for 2 min at 25 mA. The grid was blotted for 3 s at force 1 before being plunged into liquid ethane using an FEI MarkIV Vitrobot at 8 °C and 100% humidity. Micrographs were collected using an FEI Polara F30 microscope operated at 300 keV in counting mode on a K2 Summit direct electron detector (Gatan). Data were collected at a nominal magnification of ×31,000 at a defocus range of −1.4 to −3.5 μm under low dose conditions. Specimens were exposed for 10 s at ∼11 e/pix/s over 50 frames resulting in a total dose of ∼70 e/Å2 using SerialEM automated data collection (
      • Mastronarde D.N.
      Automated electron microscope tomography using robust prediction of specimen movements.
      ). The pixel size of the image was 1.247 Å.

      Cryo-EM image processing for hIP3R-3

      Motion correction and CTF parameter determination was done using the on-the-fly processing software Focus (
      • Biyani N.
      • Righetto R.D.
      • McLeod R.
      • Caujolle-Bert D.
      • Castano-Diez D.
      • Goldie K.N.
      • Stahlberg H.
      Focus: the interface between data collection and data processing in cryo-EM.
      ). All images were motion corrected from frames 0–44 using motioncor2 (
      • Zheng S.Q.
      • Palovcak E.
      • Armache J.P.
      • Verba K.A.
      • Cheng Y.
      • Agard D.A.
      MotionCor2: anisotropic correction of beam-induced motion for improved cryo-electron microscopy.
      ) with a dose weighting parameter of 1.4 e/Å/frame. CTF parameters for each image was calculated using Gctf (
      • Zhang K.
      Gctf: Real-time CTF determination and correction.
      ) and used to remove micrographs with low maximum resolution or high astigmatism. Motion-corrected images were then imported into RELION-2 (
      • Scheres S.H.
      RELION: implementation of a Bayesian approach to cryo-EM structure determination.
      ). RELION-3 was used as the new version became available (
      • Zivanov J.
      • Nakane T.
      • Forsberg B.O.
      • Kimanius D.
      • Hagen W.J.
      • Lindahl E.
      • Scheres S.H.
      New tools for automated high-resolution cryo-EM structure determination in RELION-3.
      ). Autopicking was done using representative class averages obtained from 757 manual picked particles as templates. Particles were extracted at a box size of 350 × 350 pixels and binned to 64 × 64 pixels. 2D class averages were determined using 25 iterations of classification. Particles that generated class averages showing well-defined domain structure were re-extracted at 340 × 340 pixels without binning and were subject to 2D classification. Particles that generated 2D class averages showing clear secondary structure subparticle features were subject to 3D classification. The EM density map of the IP3R-1 (EMD-6369) was scaled and clipped, using e2proc3d.py (EMAN) (
      • Tang G.
      • Peng L.
      • Baldwin P.R.
      • Mann D.S.
      • Jiang W.
      • Rees I.
      • Ludtke S.J.
      EMAN2: an extensible image processing suite for electron microscopy.
      ), to match our pixel and box size, filtered to 60 Å, and used as an initial model for 3D classification into 6 classes with no symmetry imposed. Two classes had impaired density for one of the subunits and excluded from further analysis. Refinement and reconstruction of the particles belonging to the other four classes were performed using cisTEM by imposing C4 symmetry (
      • Grant T.
      • Rohou A.
      • Grigorieff N.
      cisTEM, user-friendly software for single-particle image processing.
      ) (Table 1). The resulting map was sharpened by applying a B-factor of −90 Å2 using Rosenthal and Henderson's method (
      • Rosenthal P.B.
      • Henderson R.
      Optimal determination of particle orientation, absolute hand, and contrast loss in single-particle electron cryomicroscopy.
      ). The final average resolution at the “gold standard” 0.143 cutoff was 3.8 Å. Half-maps were generated using the 3D-generated module in cisTEM. Local resolution was also calculated in ResMap (
      • Kucukelbir A.
      • Sigworth F.J.
      • Tagare H.D.
      Quantifying the local resolution of cryo-EM density maps.
      ) (Fig. S4). Chimera (
      • Pettersen E.F.
      • Goddard T.D.
      • Huang C.C.
      • Couch G.S.
      • Greenblatt D.M.
      • Meng E.C.
      • Ferrin T.E.
      UCSF Chimera: a visualization system for exploratory research and analysis.
      ), COOT (
      • Emsley P.
      • Lohkamp B.
      • Scott W.G.
      • Cowtan K.
      Features and development of Coot.
      ) and The PyMOL Molecular Graphics System (Version 2.0, Schrödinger, LLC) were used for visualization and figure preparation.
      Table 1Cryo-EM data collection, refinement, and validation statistics
      Data collection and processing
      MicroscopeFEI Polara TF30
      DetectorGatan K2 summit
      Nominal magnification×31,000
      Voltage (kV)300
      Electron exposure (e/Å2)70
      Defocus range (μm)−1.4 to −3.5
      Pixel size (Å)1.247
      Symmetry imposedC4
      Number of micrographs3,899
      Initial particle images110,510
      Final particle images82,511
      Map resolution (Å)/FSC threshold3.8/0.143
      Refinement
      Resolution (Å)3.8
      B-factor used for map sharpening (Å2)−90
      Model composition
      Non-hydrogen atoms51,564
      Protein residues8,216
      Zinc4
      Mean B factors (Å2)
      Protein155.0
      Zinc115.7
      R.m.s. deviations
      Bond lengths (Å)0.002
      Bond angles (°)0.532
      Validation
      Molprobity score0.94
      Clashscore0.26
      Poor rotamers (%)0
      Ramachandran plot
      Favored (%)95.13
      Allowed (%)4.87
      Disallowed (%)0.0

      Symmetry expansion, partial signal subtraction, and focused 3D classification

      To resolve the unaccounted density better, we first treated each subunit as a single particle and expanded the dataset by using “relion_particle_symmetry_expand” command based on the C4 symmetry and the refined orientation parameters calculated during 3D refinement using RELION-3 for the particles. This process increased the number of particles to 330,044. We created a mask around the hIP3R-3 tetramer using MaskCreate in RELION-3 and removed the part that corresponds to the IP3R-3 NTD of one of the subunits using the “volume erase” function of Chimera. The new mask was then used to subtract the signal from the expanded particles using RELION-3 resulting in 330,044 particle images containing a signal for only the IP3R-3 NTD of one of the subunits. These particles were then subjected to 3D classification using a mask covering the IP3R-3 NTD, and orientation parameters for the particles from symmetry expansion step. At this stage, we tried several different classification strategies: 3D classification into 4, 6, or 8 classes, 3D classification using a mask excluding the ARM2, and re-classification of the 3D classes into subclasses. The results mostly agreed with each other, and we used the maps from the classification into 6 classes for further analysis (Fig. S5A). Refinement of the particles belonging to Class 5 was performed using cisTEM. The final average resolution at the gold standard 0.143 cutoff was 4.5 Å.

      Model building

      High resolution crystal structures of the β-TF1 domain of mouse IP3R-3 (PDB ID 3JRR) (
      • Chan J.
      • Yamazaki H.
      • Ishiyama N.
      • Seo M.D.
      • Mal T.K.
      • Michikawa T.
      • Mikoshiba K.
      • Ikura M.
      Structural studies of inositol 1,4,5-trisphosphate receptor: coupling ligand binding to channel gating.
      ) and β-TF2 and part of the ARM1 domain of the rat IP3R-1 (PDB ID 3UJ4)(
      • Seo M.D.
      • Velamakanni S.
      • Ishiyama N.
      • Stathopulos P.B.
      • Rossi A.M.
      • Khan S.A.
      • Dale P.
      • Li C.
      • Ames J.B.
      • Ikura M.
      • Taylor C.W.
      Structural and functional conservation of key domains in InsP3 and ryanodine receptors.
      ) were docked into the cryo-EM map followed by rigid-body fitting of the individual β-TF1, β-TF2, and ARM1 domains into the cryo-EM map using COOT (
      • Emsley P.
      • Lohkamp B.
      • Scott W.G.
      • Cowtan K.
      Features and development of Coot.
      ). The resulting model was manually modified to have the correct hIP3R-3 residue assignment and fit to the cryo-EM map. The rest of the model was built manually using COOT and refined against the cryo-EM map using Phenix real space refinement (
      • Afonine P.
      • Headd J.
      • Terwilliger T.
      • Adams P.
      New tool: phenix.real_space_refine.
      ). Residues without clear density for their side chains were built without their side chains (i.e. as alanines), while maintaining their correct labeling for the amino acid type.

      Expression and purification of soluble hIP3R-3 constructs

      The gene encoding the hIP3R-3 LBD (Met-4 to Asn-602) was subcloned into pAceBac1 vector with an N-terminal OneStrep tag followed by a tobacco etch virus protease cleavage site. This construct was modified to prepare the hIP3R-3 LBD + SBP fusion construct that encodes residues 4 to 603 followed by the Ala-Gly-Pro-Gly-Gly linker and residues 1128 to 1168. For MST experiments, the gene encoding the hIP3R-3 LBD (residues 4 to 602) and the hIP3R-3 NTD (residues 4–1799) were subcloned into pAceBac1 vector with an N-terminal OneStrep tag followed by the gene encoding eGFP and tobacco etch virus protease cleavage site. The truncation construct hIP3R-3 NTD-ΔSBP was prepared by removing residues between Gly-1132 and Gly-1156 with Asn-Gln-Ala. All constructs were incorporated into baculovirus using the Multibac expression system (
      • Fitzgerald D.J.
      • Berger P.
      • Schaffitzel C.
      • Yamada K.
      • Richmond T.J.
      • Berger I.
      Protein complex expression by using multigene baculoviral vectors.
      ). The constructs were expressed using the Sf9/Baculovirus system (DH10multibac). Sf9 cells (∼4.0 × 106 cells/ml) were harvested by centrifugation (1952 × g, 20 min) 48 h post-infection. The cell pellet was resuspended in lysis buffer composed of 200 mm NaCl, 20 mm Tris-HCl, pH 8.0, 10% glycerol (v/v), 10 mm β-ME, and 1 mm phenylmethylsulfonyl fluoride. Cells were lysed using Avestin EmulsiFlex-C3 system (greater than 10,000 p.s.i.) and centrifuged at 40,000 rpm (Ti45 rotor) for 45 min. Supernatant was recovered and incubated with Strep-XT-Superflow resin (IBA Biotagnology) for 2 h at 4 °C. The resin was then washed with the wash buffer (200 mm NaCl, 20 mm Tris-HCl, pH 8.0, 10% glycerol, and 10 mm β-ME) and the protein was eluted with the elution buffer (200 mm NaCl, 50 mm Tris, pH 8.2, 100 mmd-biotin, and 10 mm β-ME). The proteins used for ITC experiments were further purified by SEC using Superdex 200 column (GE Healthcare) equilibrated with the SEC buffer composed of 200 mm NaCl, 20 mm Tris-HCl, pH 8.0, 10% glycerol (v/v), and 0.5 mm TCEP. SEC for the proteins used in MST experiments were performed using a buffer composed of 200 mm NaCl, 20 mm Tris-HCl, pH 8.0, and 0.5 mm TCEP.

      Isothermal titration calorimetry (ITC)

      All proteins used in ITC experiments were dialyzed against the same buffer solution composed of 200 mm NaCl, 20 mm Tris-HCl, pH 8.0, 10% glycerol (v/v), and 0.5 mm TCEP to avoid the possible changes in the salt concentration and pH. IP3 solution was prepared by dissolving the powder in the dialysis solution. ITC experiments were conducted on an Auto ITC 200 instrument at 20 °C by successive injections of 3 μl of 0.8 mm IP3 to 0.08 mm of the protein solutions at 200-s intervals following an initial injection of 0.2 μl of 0.8 mm IP3. Data analysis was performed using the software Origin, version 7.0, with the MicroCal ITC analysis module.

      Microscale thermophoresis (MST)

      MST experiments were conducted on a Monolith NT.115 series (Nanotemper Technologies). Proteins were diluted to 200 nm in 1× MST buffer (50 mm Tris-HCl, 150 mm NaCl, 10 mm MgCl2, 0.05% Tween 20). IP3 stock solution (2 mm in water) was diluted to 15 μm using 1× MST buffer. A 2-fold serial dilution of IP3 in 18 NT.115 standard capillaries was prepared, with 15 μm IP3 as the highest concentration and 100 nm protein per capillary. Excitation power was set to 20% and MST laser power was set to medium. Data analysis was performed using the software MO analysis (Nanotemper Technologies).

      Data and software availability

      Cryo-EM maps and atomic coordinates have been deposited in the EMDB and PDB under the accession codes EMD-20849 (tetramer with C4 symmetry, PDB ID 6UQK) and EMD-20850 (IP3R-3 NTD, focused refinement with no imposed symmetry). All other source data are available from the corresponding authors upon request.

      Author contributions

      C. M. A., E. A. L., T. N., and E. K. conceptualization; C. M. A., E. A. L., C. J. R., and E. K. data curation; C. M. A., E. A. L., and E. K. formal analysis; C. M. A. and E. K. validation; C. M. A., E. A. L., C. J. R., T. N., and E. K. investigation; C. M. A., E. A. L., C. J. R., T. N., and E. K. writing-review and editing; T. N. and E. K. resources; T. N. and E. K. funding acquisition; T. N. and E. K. writing-original draft; T. N. and E. K. project administration; E. K. supervision.

      Acknowledgments

      EM data were collected by C. M. A. using FEI Polara and FEI TF20 at the Center for Structural Biology's Molecular Cryo-EM facility at Vanderbilt University. We thank Drs. Scott Collier and Elad Binshtein for their support at the facility. We thank Dr. Tim Grant for suggestions for data processing. We thank Drs. James Crowe and Lauren P. Jackson for kindly sharing the ITC instrument and Drs. Roger J. Colbran and Hassane Mchaourab for critically reviewing the manuscript. This work was supported in part using the CPU and GPU resources of the Advanced Computing Center for Research and Education (ACCRE) at Vanderbilt University. We used the DORS storage system supported by National Institutes of Health Grant S10RR031634 (to Jarrod Smith). We acknowledge the use of SBGrid supported software (
      • Morin A.
      • Eisenbraun B.
      • Key J.
      • Sanschagrin P.C.
      • Timony M.A.
      • Ottaviano M.
      • Sliz P.
      Collaboration gets the most out of software.
      ).

      Supplementary Material

      References

        • Foskett J.K.
        • White C.
        • Cheung K.H.
        • Mak D.O.
        Inositol trisphosphate receptor Ca2+ release channels.
        Physiol. Rev. 2007; 87 (17429043): 593-658
        • Parys J.B.
        • De Smedt H.
        Inositol 1,4,5-trisphosphate and its receptors.
        Adv. Exp. Med. Biol. 2012; 740 (22453946): 255-279
        • Berridge M.J.
        The inositol trisphosphate/calcium signaling pathway in health and disease.
        Physiol. Rev. 2016; 96 (27512009): 1261-1296
        • Mikoshiba K.
        Role of IP3 receptor signaling in cell functions and diseases.
        Adv. Biol. Regul. 2015; 57 (25497594): 217-227
        • Sato-Miyaoka M.
        • Hisatsune C.
        • Ebisui E.
        • Ogawa N.
        • Takahashi-Iwanaga H.
        • Mikoshiba K.
        Regulation of hair shedding by the type 3 IP3 receptor.
        J. Invest. Dermatol. 2012; 132 (22572820): 2137-2147
        • Miura H.
        • Nakayama A.
        • Shindo Y.
        • Kusakabe Y.
        • Tomonari H.
        • Harada S.
        Expression of gustducin overlaps with that of type III IP3 receptor in taste buds of the rat soft palate.
        Chem. Senses. 2007; 32 (17566068): 689-696
        • Hisatsune C.
        • Yasumatsu K.
        • Takahashi-Iwanaga H.
        • Ogawa N.
        • Kuroda Y.
        • Yoshida R.
        • Ninomiya Y.
        • Mikoshiba K.
        Abnormal taste perception in mice lacking the type 3 inositol 1,4,5-trisphosphate receptor.
        J. Biol. Chem. 2007; 282 (17925404): 37225-37231
        • Mendes C.C.
        • Gomes D.A.
        • Thompson M.
        • Souto N.C.
        • Goes T.S.
        • Goes A.M.
        • Rodrigues M.A.
        • Gomez M.V.
        • Nathanson M.H.
        • Leite M.F.
        The type III inositol 1,4,5-trisphosphate receptor preferentially transmits apoptotic Ca2+ signals into mitochondria.
        J. Biol. Chem. 2005; 280 (16192275): 40892-40900
        • Hayashi T.
        • Su T.P.
        Sigma-1 receptor chaperones at the ER-mitochondrion interface regulate Ca2+ signaling and cell survival.
        Cell. 2007; 131 (17981125): 596-610
        • Kuchay S.
        • Giorgi C.
        • Simoneschi D.
        • Pagan J.
        • Missiroli S.
        • Saraf A.
        • Florens L.
        • Washburn M.P.
        • Collazo-Lorduy A.
        • Castillo-Martin M.
        • Cordon-Cardo C.
        • Sebti S.M.
        • Pinton P.
        • Pagano M.
        PTEN counteracts FBXL2 to promote IP3R3- and Ca2+-mediated apoptosis limiting tumour growth.
        Nature. 2017; 546 (28614300): 554-558
        • Mound A.
        • Vautrin-Glabik A.
        • Foulon A.
        • Botia B.
        • Hague F.
        • Parys J.B.
        • Ouadid-Ahidouch H.
        • Rodat-Despoix L.
        Downregulation of type 3 inositol (1,4,5)-trisphosphate receptor decreases breast cancer cell migration through an oscillatory Ca2+ signal.
        Oncotarget. 2017; 8 (29069790): 72324-72341
        • Kang S.S.
        • Han K.S.
        • Ku B.M.
        • Lee Y.K.
        • Hong J.
        • Shin H.Y.
        • Almonte A.G.
        • Woo D.H.
        • Brat D.J.
        • Hwang E.M.
        • Yoo S.H.
        • Chung C.K.
        • Park S.H.
        • Paek S.H.
        • Roh E.J.
        • et al.
        Caffeine-mediated inhibition of calcium release channel inositol 1,4,5-trisphosphate receptor subtype 3 blocks glioblastoma invasion and extends survival.
        Cancer Res. 2010; 70 (20103623): 1173-1183
        • Sakakura C.
        • Hagiwara A.
        • Fukuda K.
        • Shimomura K.
        • Takagi T.
        • Kin S.
        • Nakase Y.
        • Fujiyama J.
        • Mikoshiba K.
        • Okazaki Y.
        • Yamagishi H.
        Possible involvement of inositol 1,4,5-trisphosphate receptor type 3 (IP3R3) in the peritoneal dissemination of gastric cancers.
        Anticancer Res. 2003; 23 (14666665): 3691-3697
        • Shibao K.
        • Fiedler M.J.
        • Nagata J.
        • Minagawa N.
        • Hirata K.
        • Nakayama Y.
        • Iwakiri Y.
        • Nathanson M.H.
        • Yamaguchi K.
        The type III inositol 1,4,5-trisphosphate receptor is associated with aggressiveness of colorectal carcinoma.
        Cell Calcium. 2010; 48 (21075448): 315-323
        • Bononi A.
        • Giorgi C.
        • Patergnani S.
        • Larson D.
        • Verbruggen K.
        • Tanji M.
        • Pellegrini L.
        • Signorato V.
        • Olivetto F.
        • Pastorino S.
        • Nasu M.
        • Napolitano A.
        • Gaudino G.
        • Morris P.
        • Sakamoto G.
        • et al.
        BAP1 regulates IP3R3-mediated Ca2+ flux to mitochondria suppressing cell transformation.
        Nature. 2017; 546 (28614305): 549-553
        • Giorgi C.
        • Ito K.
        • Lin H.K.
        • Santangelo C.
        • Wieckowski M.R.
        • Lebiedzinska M.
        • Bononi A.
        • Bonora M.
        • Duszynski J.
        • Bernardi R.
        • Rizzuto R.
        • Tacchetti C.
        • Pinton P.
        • Pandolfi P.P.
        PML regulates apoptosis at endoplasmic reticulum by modulating calcium release.
        Science. 2010; 330 (21030605): 1247-1251
        • Szado T.
        • Vanderheyden V.
        • Parys J.B.
        • De Smedt H.
        • Rietdorf K.
        • Kotelevets L.
        • Chastre E.
        • Khan F.
        • Landegren U.
        • Söderberg O.
        • Bootman M.D.
        • Roderick H.L.
        Phosphorylation of inositol 1,4,5-trisphosphate receptors by protein kinase B/Akt inhibits Ca2+ release and apoptosis.
        Proc. Natl. Acad. Sci. U.S.A. 2008; 105 (18250332): 2427-2432
        • Fan G.
        • Baker M.L.
        • Wang Z.
        • Baker M.R.
        • Sinyagovskiy P.A.
        • Chiu W.
        • Ludtke S.J.
        • Serysheva I.I.
        Gating machinery of InsP3R channels revealed by electron cryomicroscopy.
        Nature. 2015; 527 (26458101): 336-341
        • da Fonseca P.C.
        • Morris S.A.
        • Nerou E.P.
        • Taylor C.W.
        • Morris E.P.
        Domain organization of the type 1 inositol 1,4,5-trisphosphate receptor as revealed by single-particle analysis.
        Proc. Natl. Acad. Sci. U.S.A. 2003; 100 (12651956): 3936-3941
        • Hamada K.
        • Terauchi A.
        • Mikoshiba K.
        Three-dimensional rearrangements within inositol 1,4,5-trisphosphate receptor by calcium.
        J. Biol. Chem. 2003; 278 (14593123): 52881-52889
        • Jiang Q.X.
        • Thrower E.C.
        • Chester D.W.
        • Ehrlich B.E.
        • Sigworth F.J.
        Three-dimensional structure of the type 1 inositol 1,4,5-trisphosphate receptor at 24 Å resolution.
        EMBO J. 2002; 21 (12110570): 3575-3581
        • Sato C.
        • Hamada K.
        • Ogura T.
        • Miyazawa A.
        • Iwasaki K.
        • Hiroaki Y.
        • Tani K.
        • Terauchi A.
        • Fujiyoshi Y.
        • Mikoshiba K.
        Inositol 1,4,5-trisphosphate receptor contains multiple cavities and L-shaped ligand-binding domains.
        J. Mol. Biol. 2004; 336 (14741211): 155-164
        • Serysheva I.I.
        • Bare D.J.
        • Ludtke S.J.
        • Kettlun C.S.
        • Chiu W.
        • Mignery G.A.
        Structure of the type 1 inositol 1,4,5-trisphosphate receptor revealed by electron cryomicroscopy.
        J. Biol. Chem. 2003; 278 (12714606): 21319-21322
        • Nakade S.
        • Rhee S.K.
        • Hamanaka H.
        • Mikoshiba K.
        Cyclic AMP-dependent phosphorylation of an immunoaffinity-purified homotetrameric inositol 1,4,5-trisphosphate receptor (type I) increases Ca2+ flux in reconstituted lipid vesicles.
        J. Biol. Chem. 1994; 269 (8120033): 6735-6742
        • Supattapone S.
        • Worley P.F.
        • Baraban J.M.
        • Snyder S.H.
        Solubilization, purification, and characterization of an inositol trisphosphate receptor.
        J. Biol. Chem. 1988; 263 (2826483): 1530-1534
        • Fan G.
        • Baker M.R.
        • Wang Z.
        • Seryshev A.B.
        • Ludtke S.J.
        • Baker M.L.
        • Serysheva I.I.
        Cryo-EM reveals ligand induced allostery underlying InsP3R channel gating.
        Cell Res. 2018; 28 (30470765): 1158-1170
        • Paknejad N.
        • Hite R.K.
        Structural basis for the regulation of inositol trisphosphate receptors by Ca(2+) and IP3.
        Nat. Struct. Mol. Biol. 2018; 25 (30013099): 660-668
        • Chan J.
        • Yamazaki H.
        • Ishiyama N.
        • Seo M.D.
        • Mal T.K.
        • Michikawa T.
        • Mikoshiba K.
        • Ikura M.
        Structural studies of inositol 1,4,5-trisphosphate receptor: coupling ligand binding to channel gating.
        J. Biol. Chem. 2010; 285 (20843799): 36092-36099
        • Seo M.D.
        • Velamakanni S.
        • Ishiyama N.
        • Stathopulos P.B.
        • Rossi A.M.
        • Khan S.A.
        • Dale P.
        • Li C.
        • Ames J.B.
        • Ikura M.
        • Taylor C.W.
        Structural and functional conservation of key domains in InsP3 and ryanodine receptors.
        Nature. 2012; 483 (22286060): 108-112
        • Scheres S.H.W.
        Processing of Structurally Heterogeneous Cryo-EM Data in RELION.
        Method Enzymol. 2016; 579: 125-157
        • Yang C.
        • Zhang S.
        • He P.
        • Wang C.
        • Huang J.
        • Zhou P.
        Self-binding peptides: folding or binding?.
        J. Chem. Inf. Model. 2015; 55 (25643174): 329-342
        • des Georges A.
        • Clarke O.B.
        • Zalk R.
        • Yuan Q.
        • Condon K.J.
        • Grassucci R.A.
        • Hendrickson W.A.
        • Marks A.R.
        • Frank J.
        Structural basis for gating and activation of RyR1.
        Cell. 2016; 167 (27662087): 145-157.e17
        • McGoldrick L.L.
        • Singh A.K.
        • Saotome K.
        • Yelshanskaya M.V.
        • Twomey E.C.
        • Grassucci R.A.
        • Sobolevsky A.I.
        Opening of the human epithelial calcium channel TRPV6.
        Nature. 2018; 553 (29258289): 233-237
        • Kalinina J.
        • Dutta K.
        • Ilghari D.
        • Beenken A.
        • Goetz R.
        • Eliseenkova A.V.
        • Cowburn D.
        • Mohammadi M.
        The alternatively spliced acid box region plays a key role in FGF receptor autoinhibition.
        Structure. 2012; 20 (22244757): 77-88
        • Iwai M.
        • Michikawa T.
        • Bosanac I.
        • Ikura M.
        • Mikoshiba K.
        Molecular basis of the isoform-specific ligand-binding affinity of inositol 1,4,5-trisphosphate receptors.
        J. Biol. Chem. 2007; 282 (17327232): 12755-12764
        • Strausberg R.L.
        • Feingold E.A.
        • Grouse L.H.
        • Derge J.G.
        • Klausner R.D.
        • Collins F.S.
        • Wagner L.
        • Shenmen C.M.
        • Schuler G.D.
        • Altschul S.F.
        • Zeeberg B.
        • Buetow K.H.
        • Schaefer C.F.
        • Bhat N.K.
        • Hopkins R.F.
        • et al.
        Generation and initial analysis of more than 15,000 full-length human and mouse cDNA sequences.
        Proc. Natl. Acad. Sci. U.S.A. 2002; 99 (12477932): 16899-16903
        • Fitzgerald D.J.
        • Berger P.
        • Schaffitzel C.
        • Yamada K.
        • Richmond T.J.
        • Berger I.
        Protein complex expression by using multigene baculoviral vectors.
        Nat. Methods. 2006; 3 (17117155): 1021-1032
        • Mastronarde D.N.
        Automated electron microscope tomography using robust prediction of specimen movements.
        J. Struct. Biol. 2005; 152 (16182563): 36-51
        • Scheres S.H.
        RELION: implementation of a Bayesian approach to cryo-EM structure determination.
        J. Struct. Biol. 2012; 180 (23000701): 519-530
        • Biyani N.
        • Righetto R.D.
        • McLeod R.
        • Caujolle-Bert D.
        • Castano-Diez D.
        • Goldie K.N.
        • Stahlberg H.
        Focus: the interface between data collection and data processing in cryo-EM.
        J. Struct. Biol. 2017; 198 (28344036): 124-133
        • Zheng S.Q.
        • Palovcak E.
        • Armache J.P.
        • Verba K.A.
        • Cheng Y.
        • Agard D.A.
        MotionCor2: anisotropic correction of beam-induced motion for improved cryo-electron microscopy.
        Nat. Methods. 2017; 14 (28250466): 331-332
        • Zhang K.
        Gctf: Real-time CTF determination and correction.
        J. Struct. Biol. 2016; 193 (26592709): 1-12
        • Zivanov J.
        • Nakane T.
        • Forsberg B.O.
        • Kimanius D.
        • Hagen W.J.
        • Lindahl E.
        • Scheres S.H.
        New tools for automated high-resolution cryo-EM structure determination in RELION-3.
        Elife. 2018; 7 (30412051): e42166
        • Tang G.
        • Peng L.
        • Baldwin P.R.
        • Mann D.S.
        • Jiang W.
        • Rees I.
        • Ludtke S.J.
        EMAN2: an extensible image processing suite for electron microscopy.
        J. Struct. Biol. 2007; 157 (16859925): 38-46
        • Grant T.
        • Rohou A.
        • Grigorieff N.
        cisTEM, user-friendly software for single-particle image processing.
        Elife. 2018; 7 (29513216): e35383
        • Rosenthal P.B.
        • Henderson R.
        Optimal determination of particle orientation, absolute hand, and contrast loss in single-particle electron cryomicroscopy.
        J. Mol. Biol. 2003; 333 (14568533): 721-745
        • Kucukelbir A.
        • Sigworth F.J.
        • Tagare H.D.
        Quantifying the local resolution of cryo-EM density maps.
        Nat. Methods. 2014; 11 (24213166): 63-65
        • Pettersen E.F.
        • Goddard T.D.
        • Huang C.C.
        • Couch G.S.
        • Greenblatt D.M.
        • Meng E.C.
        • Ferrin T.E.
        UCSF Chimera: a visualization system for exploratory research and analysis.
        J. Comput. Chem. 2004; 25 (15264254): 1605-1612
        • Emsley P.
        • Lohkamp B.
        • Scott W.G.
        • Cowtan K.
        Features and development of Coot.
        Acta Crystallogr. D Biol. Crystallogr. 2010; 66 (20383002): 486-501
        • Afonine P.
        • Headd J.
        • Terwilliger T.
        • Adams P.
        New tool: phenix.real_space_refine.
        Comput. Crystallogr. Newsletter. 2013; 4: 43-44
        • Morin A.
        • Eisenbraun B.
        • Key J.
        • Sanschagrin P.C.
        • Timony M.A.
        • Ottaviano M.
        • Sliz P.
        Collaboration gets the most out of software.
        Elife. 2013; 2 (24040512): e01456
        • Hagemans D.
        • van Belzen I.A.
        • Morán Luengo T.
        • Rudiger S.G.
        A script to highlight hydrophobicity and charge on protein surfaces.
        Front. Mol. Biosci. 2015; 2 (26528483): 56
        • Smart O.S.
        • Neduvelil J.G.
        • Wang X.
        • Wallace B.A.
        • Sansom M.S.
        HOLE: a program for the analysis of the pore dimensions of ion channel structural models.
        J. Mol. Graph. 1996; 14 (376 9195488): 354-360
        • Marcus Y.
        Ionic radii in aqueous solutions.
        Chem. Rev. 1988; 88: 1475-1498