Molecular Mechanism of pH-Induced Protrusion Configuration Switching in Piscine Betanodavirus Implies a Novel Antiviral StrategyClick to copy article linkArticle link copied!
- Petra ŠtěrbováPetra ŠtěrbováChemical Biology and Molecular Biophysics Program, Taiwan International Graduate Program, Academia Sinica, Taipei 11529, TaiwanCollege of Life Science, National Tsing Hua University, Hsinchu 30044, TaiwanInstitute of Chemistry, Academia Sinica, Taipei 11529, TaiwanMore by Petra Štěrbová
- Chun-Hsiung WangChun-Hsiung WangInstitute of Chemistry, Academia Sinica, Taipei 11529, TaiwanMore by Chun-Hsiung Wang
- Kathleen J. D. CarilloKathleen J. D. CarilloInstitute of Chemistry, Academia Sinica, Taipei 11529, TaiwanMore by Kathleen J. D. Carillo
- Yuan-Chao LouYuan-Chao LouBiomedical Translation Research Center, Academia Sinica, Taipei 11529, TaiwanMore by Yuan-Chao Lou
- Takayuki KatoTakayuki KatoGraduate School of Frontier Biosciences, Osaka University, 1-3 Yamadaoka, Suita, Osaka 565-0871, JapanMore by Takayuki Kato
- Keiichi NambaKeiichi NambaGraduate School of Frontier Biosciences, Osaka University, 1-3 Yamadaoka, Suita, Osaka 565-0871, JapanMore by Keiichi Namba
- Der-Lii M. Tzou
- Wei-Hau Chang*Wei-Hau Chang*Email: [email protected]Chemical Biology and Molecular Biophysics Program, Taiwan International Graduate Program, Academia Sinica, Taipei 11529, TaiwanInstitute of Chemistry, Academia Sinica, Taipei 11529, TaiwanGenomics Research Center, Academia Sinica, Taipei 11529, TaiwanInstitute of Physics, Academia Sinica, Taipei 11529, TaiwanMore by Wei-Hau Chang
Abstract
Many viruses contain surface spikes or protrusions that are essential for virus entry. These surface structures can thereby be targeted by antiviral drugs to treat viral infections. Nervous necrosis virus (NNV), a simple nonenveloped virus in the genus of betanodavirus, infects fish and damages aquaculture worldwide. NNV has 60 conspicuous surface protrusions, each comprising three protrusion domains (P-domain) of its capsid protein. NNV uses protrusions to bind to common receptors of sialic acids on the host cell surface to initiate its entry via the endocytic pathway. However, structural alterations of NNV in response to acidic conditions encountered during this pathway remain unknown, while detailed interactions of protrusions with receptors are unclear. Here, we used cryo-EM to discover that Grouper NNV protrusions undergo low-pH-induced compaction and resting. NMR and molecular dynamics (MD) simulations were employed to probe the atomic details. A solution structure of the P-domain at pH 7.0 revealed a long flexible loop (amino acids 311–330) and a pocket outlined by this loop. Molecular docking analysis showed that the N-terminal moiety of sialic acid inserted into this pocket to interact with conserved residues inside. MD simulations demonstrated that part of this loop converted to a β-strand under acidic conditions, allowing for P-domain trimerization and compaction. Additionally, a low-pH-favored conformation is attained for the linker connecting the P-domain to the NNV shell, conferring resting protrusions. Our findings uncover novel pH-dependent conformational switching mechanisms underlying NNV protrusion dynamics potentially utilized for facilitating NNV entry, providing new structural insights into complex NNV-host interactions with the identification of putative druggable hotspots on the protrusion.
This publication is licensed under
License Summary*
You are free to share(copy and redistribute) this article in any medium or format and to adapt(remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
License Summary*
You are free to share(copy and redistribute) this article in any medium or format and to adapt(remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
License Summary*
You are free to share(copy and redistribute) this article in any medium or format and to adapt(remix, transform, and build upon) the material for any purpose, even commercially within the parameters below:
Creative Commons (CC): This is a Creative Commons license.
Attribution (BY): Credit must be given to the creator.
*Disclaimer
This summary highlights only some of the key features and terms of the actual license. It is not a license and has no legal value. Carefully review the actual license before using these materials.
Results
Cryo-EM Reveals That GNNV Protrusions Are Dynamic Entities
Figure 1
Figure 1. GNNV atomic model and cryo-EM structures in three different pH environments. (A) Atomic models of GNNV S-domain (PDB 4WIZ) and P-domain (PDB 4RFU). (B) Surface views of GNNV virus-like particles at pH 8.0 (left panel), 6.5 (middle panel), and 5.0 (right panel). The views are colored with a heatmap according to the local radius. (C) Conformational changes of the protrusion in response to the change in pH: 8.0 (left panel), 6.5 (middle panel), and 5.0 (right panel). The protrusions rotate clockwise and shift ∼5 Å toward the capsid shell as the pH decreases. (D) Enlarged views of pores (at 5- and 3-fold axis) in the capsid shell surface for different pH environments. The size of a pore at each pH condition is indicated. To highlight the pore at the 3-fold axis, the densities underneath are colored in light purple.
NMR Spectra Indicate That the P-Domain of GNNV Contains Highly pH-Sensitive Regions
Figure 2
Figure 2. Effects of pH on GNNV-P in solution. (A) 1H–15N HSQC spectra of GNNV-P recorded at pH 7.0 (blue) and 5.0 (red). (B) Small box in panel (A) is enlarged. The six peaks, in green, at pH 7.0 are from Y279, L263, I222, I323, Q322, and R321, and the peaks, in red, at pH 5.0 are from E217 and I222. Y279, L263, I323, and Q322 disappear in the 1H–15N HSQC as the pH decreases. (C) Residue chemical shift perturbations (CSPs) by comparing 15N-labeled (pH 7.0) and 13C-, 15N-, and D-labeled (pH 5.0) HSQC spectra. The pH-sensitive Regions I (L263–Y279), II (W301–N303), and III (V312–V327) are highlighted in orange, yellow, and green, respectively. Residues with CSPs >2 standard deviations in pH-sensitive Regions I and III are highlighted in red and dark green, respectively. (D) pH-sensitive residues presented in panel (C) are mapped onto a surface representation of the GNNV-P crystal structure, (22) with the same color scheme as in panel (B).
Determination of the Structure of GNNV-P in Solution at pH 7.0 Reveals That aa 311–330 Form a Long Flexible Loop
Figure 3
Figure 3. Structure of GNNV-P determined in solution at neutral pH. (A) Sedimentation velocity analytical ultracentrifugation (SV AUC) data of GNNV-P obtained at pH 7.0 (black), 6.0 (blue), and 5.0 (red) fitted to a continuous sedimentation coefficient distribution c(s) model. Sedimentation coefficients and the molecular weight determined by SEDFIT are denoted above each peak. (B) Backbone (N, Cα, and C′) superimposition of the ensemble of 20 low-energy conformations. The N-terminal is colored blue, and the C-terminal is colored red. (C) Cartoon representation of the GNNV-P solution structure with β-strands and α-helices highlighted in red and blue, respectively. (D) Superimposition of the GNNV-P structure determined by NMR at neutral pH (pink) and by the X-ray crystal (PDB 4RFU, blue). (E) Schematic representation of GNNV-P secondary structures, as determined by NMR (pH 7.0) and X-ray crystallography (pH 6.5). β-strands are displayed as black arrows and α-helices as gray barrels, with bordering residues indicated.
MD Simulations Predict That GNNV-P Forms a Trimer at pH 5.0 with Critical Subunit Interactions Perturbed by Site-Directed Mutagenesis
Figure 4
Figure 4. MD-predicted structure of the GNNV-P oligomer at low pH. (A) Top and side view cartoon representations of the GNNV-P trimer formed under acidic pH conditions, as determined by molecular dynamics (MD) simulations. Chain A (blue), chain B (pink), and chain C (green). (B) Intermolecular interactions between neighboring GNNV-P units in the trimer. Residues at the trimeric interface are depicted as yellow sticks, and interactions are shown as dashed magenta lines. (C–E) Details of trimeric interface interactions between GNNV-P chains A and B, as boxed in panel (B).
residue | intermolecular interactions at pH 5.0 | mutation | AUC results (pH 5.0) |
---|---|---|---|
Region I | |||
R276 | H-bond (Y315) | R276A | oligomer |
W280 | CH–π (P326) | W280A | monomer |
C–π (P326) | |||
hydrophobic (L324) | |||
H281 | H281Y | oligomer | |
Region II | |||
W301a | H-bond (Q225) | W301A | N/A |
Region III | |||
Q322 | H-bond (Y315) | Q322A | monomer |
hydrophobic (I300) | |||
I323 | I323A | oligomer | |
L324 | hydrophobic (W280) | L324A | monomer |
P326 | CH–π (W280) | P326A | monomer |
C–π (W280) |
Mutation resulted in a misfolded protein.
NMR Signatures of GNNV-P Conformational Change at Low pH
Figure 5
Figure 5. GNNV-P undergoes a low-pH-induced conformational change. (A) Secondary structure propensities of GNNV-P at pH 7.0 and 5.0, calculated using secondary chemical ΔCα–ΔCβ shifts and plotted against the amino acid sequence. The Q320–P324 region showing an increased β-strand propensity at pH 5.0 is highlighted in gray. The secondary structures are shown above each chart, with arrows and cylinders representing β-strands and α-helices, respectively. (B) MD simulations revealed the formation of a short β-strand (comprising residues Q322–L324) within the F′–G′ loop at low pH. Hydrogen bonds between R321–L324 and D266–S264 were identified as stabilizing this region. (C) Superimposition of GNNV-P at pH 7.0 (NMR structure, blue) and 5.0 (MD model, red), showing the open and closed conformations of the F′–G′ loop, respectively. (D) 1H–15N HSQC spectra of uniform U–D-, 13C-, and 15N-labeled GNNV-P at pH 5.0. The regions with duplicate signals for residues 216–220 in the linker region are marked with black boxes. (E–G) Details of the uniform U–D-, 13C-, and 15N-labeled GNNV-P 1H–15N HSQC spectra highlighted by black boxes in panel (D). (H) Sequence of the N-terminal linker T214-P221 and a stick representation of the linker region with proline residues colored red and residues presenting duplicate NMR signals colored blue. (I) Schematic representation of cis–trans isomerization of an Xaa–Pro peptide bond. Proline residues can switch between the trans (blue) and cis (red) conformations.
Interactions of GNNV-P with Host Surface Glycan Receptors
Figure 6
Figure 6. Model of Neu5Ac-Lac binding to GNNV-P at neutral pH. (A) Surface representation of the GNNV-P monomer in complex with Neu5Ac-(α2,3)-Lac (green) and Neu5Ac-(α2,3)-Lac (yellow). (B) Details of Neu5Ac-(α2,3)-Lac binding to GNNV-P in the open pocket conformation. (C) Details of Neu5Ac-(α2,6)-Lac binding to GNNV-P in the open pocket conformation. In panels (B) and (C), residues interacting with Neu5Ac-Lac are shown as sticks, and selected contacts between GNNV-P and Neu5Ac-Lac are represented by pink dashed lines.
Discussion
Advancement from the Crystal Structure─GNNV Structure Configuration for Cell Attachment
Low-pH-Induced P-Domain Structural Transitions and Self-Assembly
Deep Pocket in the P-Domain and Its Implications
Figure 7
Figure 7. Model of GNNV interactions with host cell surface receptors. During infection, GNNV attaches to cells by interacting with the sialic acid moiety on the host cell surface and with cellular receptors (HSP70, HSP90ab1, or nectin-4). At weakly alkaline or neutral pH for host cell entry, protrusions on the GNNV surface are in extended and loose configuration, with the sialic acid binding pocket open, and linker region (red) and host-determining region (blue) accessible for interactions with cellular receptors. Acidification in late endosomes induces conformational change of the GNNV P-domain to result in resting/prone and compact protrusions. In this configuration, the sialic acid binding pocket is closed so that sialic acids are withdrawn to the tip of the protrusions. In addition, the accessibility of the linker region to a cellular receptor is also reprogrammed, leading to its potential GNNV detachment for endosomal escape. Note that the potential of the protrusion tip in directly interacting with the endosomal membrane may be altered by the pH-induced modification of the protrusion surface hydrophobicity distribution (Figure S20).
Malleable Linker That Connects the P-Domain to the S-Domain
Methods
Sample Preparation of GNNV and VLPs for Cryo-EM
Cryo-EM Data Acquisition
Single-Particle Image Processing and 3D Reconstruction
Protein Construct and Site-Directed Mutagenesis
Protein Expression and Purification
Sedimentation Velocity Analytical Ultracentrifugation (SV AUC)
NMR Spectroscopy and Structure Determination
Amide Hydrogen–Deuterium Exchange Rate (HXD)
Chemical Shift Perturbation and Secondary Chemical Shifts Calculation
Molecular Dynamics (MD) Simulation
Molecular Docking Analysis
SV-AUC Data Interpretation and Protrusion Hydrophobicity Analysis
Data Availability
Cryo-EM maps of GNNV VLP at pH 8.0, 6.5, and 5.0 are available from the EMDB data bank with accession numbers EMDB-39212, 39213, and 39214, and the corresponding atomic models are with PDB entries of 8YF6, 8YF7, and 8YF8; Cryo-EM maps of the GNNV virion at pH 6.5 and 5.0 have EMDB accession numbers 39215 and 39217, and the atomic model of the GNNV virion at pH 6.5 is with PDB entry 8YF9. 1H, 13C, and 15N chemical shift assignments of GNNV-P at pH 5.0 have been deposited to the Biological Magnetic Resonance Bank (BMRB) with Accession No. 52218. The NMR structure of the GNNV P-domain has been deposited to the Protein Data Bank (PDB entry 8XID).
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsinfecdis.4c00407.
Experimental methods; Figures S1–S21; Tables S1 and S2 (PDF)
Conformational change of GNNV VLP from pH 8.0 to 5.0 (Movie S1) (MP4)
Enlarged views of GNNV protrusion conformational change from pH 8.0 to 5.0 (Movie S2) (MP4)
Hypothetical atomic model of GNNV-P conformational change from pH 8.0 to 5.0 (Movie S3) (MP4)
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.
Acknowledgments
The authors are indebted to Yeukuang Hwu and Yi-Yun Chen at the Institute of Physics, Academia Sinica, for supporting the use of JEM-2100F (JEOL Ltd., Akishima, Tokyo, Japan), as well as to Dong-Hua Chen, David Bushnell, and Roger Kornberg at the Stanford University for supporting the use of Technai F20 (FEI, Hillsboro, OR, USA). The operation of cryo-ARM at the Osaka University was supported by a Japan Joint Research Grant to K.N. Note that all of the cryo-EM data for this work were obtained in the period of 2013–2017 prior to installation of the Academia Sinica Cryo-EM Core Facility (ASCEM) in 2019 supported by [AS-CFII-108-110]. The authors thank the Academia Sinica High-Field NMR Center (HFNMRC; funded by Academia Sinica Core Facility and Innovative Instrument Project AS-CFII-111-214) for technical support. The authors thank the Medicinal Chemistry and Analytical Core Facilities, funded by Academia Sinica Core Facility and Innovative Instrument Project AS-NBRPCF-111-201, for technical support with NMR data. We thank Mr. Kun Hung Chen in the Biophysics Core Facility, funded by Academia Sinica Core Facility and Innovative Instrument Project AS-CFII-111-201, for performing the sedimentation velocity analytical ultracentrifugation experiments. The authors thank the DNA Sequencing Core Facility of the Institute of Biomedical Sciences, Academia Sinica (funded by Academia Sinica Core Facility and Innovative Instrument Project AS-CFII-111-211), for DNA sequencing analysis. This study made use of NMRbox: National Center for Biomolecular NMR Data Processing and Analysis, a Biomedical Technology Research Resource (BTRR), supported by NIH Grant P41GM111135 (NIGMS).
References
This article references 86 other publications.
- 1Baker, R. E.; Mahmud, A. S.; Miller, I. F.; Rajeev, M.; Rasambainarivo, F.; Rice, B. L.; Takahashi, S.; Tatem, A. J.; Wagner, C. E.; Wang, L.-F. Infectious disease in an era of global change. Nat. Rev. Microbiol. 2022, 20 (4), 193– 205, DOI: 10.1038/s41579-021-00639-zGoogle Scholar1https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXit1Ciu77L&md5=e91900ca47ac5d7cf14a480ce6182ef9Infectious disease in an era of global changeBaker, Rachel E.; Mahmud, Ayesha S.; Miller, Ian F.; Rajeev, Malavika; Rasambainarivo, Fidisoa; Rice, Benjamin L.; Takahashi, Saki; Tatem, Andrew J.; Wagner, Caroline E.; Wang, Lin-Fa; Wesolowski, Amy; Metcalf, C. Jessica E.Nature Reviews Microbiology (2022), 20 (4), 193-205CODEN: NRMACK; ISSN:1740-1526. (Nature Portfolio)A review on. The twenty-first century has witnessed a wave of severe infectious disease outbreaks, not least the COVID-19 pandemic, which has had a devastating impact on lives and livelihoods around the globe. The 2003 severe acute respiratory syndrome coronavirus outbreak, the 2009 swine flu pandemic, the 2012 Middle East respiratory syndrome coronavirus outbreak, the 2013-2016 Ebola virus disease epidemic in West Africa and the 2015 Zika virus disease epidemic all resulted in substantial morbidity and mortality while spreading across borders to infect people in multiple countries. At the same time, the past few decades have ushered in an unprecedented era of technol., demog. and climatic change: airline flights have doubled since 2000, since 2007 more people live in urban areas than rural areas, population nos. continue to climb and climate change presents an escalating threat to society. In this Review, we consider the extent to which these recent global changes have increased the risk of infectious disease outbreaks, even as improved sanitation and access to health care have resulted in considerable progress worldwide.
- 2Doan, Q. K.; Vandeputte, M.; Chatain, B.; Morin, T.; Allal, F. Viral encephalopathy and retinopathy in aquaculture: a review. J. Fish Dis. 2017, 40 (5), 717– 742, DOI: 10.1111/jfd.12541Google ScholarThere is no corresponding record for this reference.
- 3Cann, A. J. Replication of Viruses. In Encyclopedia of Virology (Third ed.); Mahy, B. W. J.; Van Regenmortel, M. H. V., Eds.; Academic Press, 2008; pp 406– 412.Google ScholarThere is no corresponding record for this reference.
- 4Bruinsma, R. F.; Wuite, G. J. L.; Roos, W. H. Physics of viral dynamics. Nat. Rev. Phys. 2021, 3 (2), 76– 91, DOI: 10.1038/s42254-020-00267-1Google ScholarThere is no corresponding record for this reference.
- 5Sherman, M. B.; Smith, H. Q.; Smith, T. J. The Dynamic Life of Virus Capsids. Viruses 2020, 12 (6), 618, DOI: 10.3390/v12060618Google ScholarThere is no corresponding record for this reference.
- 6Zhang, Y.; Cui, Y.; Sun, J.; Zhou, Z. H. Multiple conformations of trimeric spikes visualized on a non-enveloped virus. Nat. Commun. 2022, 13 (1), 550 DOI: 10.1038/s41467-022-28114-0Google ScholarThere is no corresponding record for this reference.
- 7Lynch, D. L.; Pavlova, A.; Fan, Z.; Gumbart, J. C. Understanding Virus Structure and Dynamics through Molecular Simulations. J. Chem. Theory Comput. 2023, 19 (11), 3025– 3036, DOI: 10.1021/acs.jctc.3c00116Google ScholarThere is no corresponding record for this reference.
- 8Crane, M.; Hyatt, A. Viruses of fish: an overview of significant pathogens. Viruses 2011, 3 (11), 2025– 2046, DOI: 10.3390/v3112025Google ScholarThere is no corresponding record for this reference.
- 9Mori, K.-I.; Nakai, T.; Muroga, K.; Arimoto, M.; Mushiake, K.; Furusawa, I. Properties of a new virus belonging to nodaviridae found in larval striped jack (Pseudocaranx dentex) with nervous necrosis. Virology 1992, 187 (1), 368– 371, DOI: 10.1016/0042-6822(92)90329-NGoogle Scholar9https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADyaK387kt1Ghug%253D%253D&md5=2e07e4427b2baf88353ad92b9de7012dProperties of a new virus belonging to nodaviridae found in larval striped jack (Pseudocaranx dentex) with nervous necrosisMori K; Nakai T; Muroga K; Arimoto M; Mushiake K; Furusawa IVirology (1992), 187 (1), 368-71 ISSN:0042-6822.Spherical virus particles were purified from larval striped jack (Pseudocaranx dentex) with nervous necrosis. The virus consists of nonenveloped particles, about 25 nm in diameter, and contains two single-stranded, positive-sense RNA molecules with molecular weights of 1.01 x 10(6)Da (RNA 1) and 0.49 x 10(6)Da (RNA 2), respectively. The RNAs do not have poly(A) sequences at the 3' terminus. Virus structural proteins consist of two proteins with molecular weights of 42 and 40 kDa. When translated into cell-free extracts of rabbit reticulocytes, RNA 1 directed the synthesis of the 1a protein (100 kDa), whereas RNA 2 synthesized the 2a protein (42 kDa), which is probably the coat protein of the virus, and a polypeptide of 40 kDa which appears to be the processed form of the 42-kDa protein. Under electron microscopic observation, the virus particles were found in the tissues of the central nervous system of the affected larval striped jack. From morphological and biochemical properties of the virus, we identified this virus as a new member of the family of Nodaviridae and designated it striped jack nervous necrosis virus.
- 10Bandín, I.; Souto, S. Betanodavirus and VER Disease: A 30-year Research Review. Pathogens 2020, 9 (2), 106, DOI: 10.3390/pathogens9020106Google ScholarThere is no corresponding record for this reference.
- 11Venter, P. A.; Schneemann, A. Nodaviruses. In Encyclopedia of Virology (Third ed.); Mahy, B. W. J.; Van Regenmortel, M. H. V., Eds.; Academic Press, 2008; pp 430– 438.Google ScholarThere is no corresponding record for this reference.
- 12Tang, L.; Lin, C.-S.; Krishna, N. K.; Yeager, M.; Schneemann, A.; Johnson, J. E. Virus-like particles of a fish nodavirus display a capsid subunit domain organization different from that of insect nodaviruses. J. Virol. 2002, 76 (12), 6370– 6375, DOI: 10.1128/JVI.76.12.6370-6375.2002Google ScholarThere is no corresponding record for this reference.
- 13Iwamoto, T.; Mise, K.; Takeda, A.; Okinaka, Y.; Mori, K.-I.; Arimoto, M.; Okuno, T.; Nakai, T. Characterization of Striped jack nervous necrosis virus subgenomic RNA3 and biological activities of its encoded protein B2. J. Gen. Virol. 2005, 86, 2807– 2816, DOI: 10.1099/vir.0.80902-0Google ScholarThere is no corresponding record for this reference.
- 14Liu, W.; Hsu, C.-H.; Hong, Y.-R.; Wu, S.-C.; Wang, C.-H.; Wu, Y.-M.; Chao, C.-B.; Lin, C.-S. Early endocytosis pathways in SSN-1 cells infected by dragon grouper nervous necrosis virus. J. Gen. Virol. 2005, 86 (9), 2553– 2561, DOI: 10.1099/vir.0.81021-0Google ScholarThere is no corresponding record for this reference.
- 15Adachi, K.; Ichinose, T.; Takizawa, N.; Watanabe, K.; Kitazato, K.; Kobayashi, N. Inhibition of betanodavirus infection by inhibitors of endosomal acidification. Arch. Virol. 2007, 152 (12), 2217– 2224, DOI: 10.1007/s00705-007-1061-7Google ScholarThere is no corresponding record for this reference.
- 16Huang, R.; Zhu, G.; Zhang, J.; Lai, Y.; Xu, Y.; He, J.; Xie, J. Betanodavirus-like particles enter host cells via clathrin-mediated endocytosis in a cholesterol-, pH- and cytoskeleton-dependent manner. Vet. Res. 2017, 48 (1), 8 DOI: 10.1186/s13567-017-0412-yGoogle ScholarThere is no corresponding record for this reference.
- 17Kumar, C. S.; Dey, D.; Ghosh, S.; Banerjee, M. Breach: Host Membrane Penetration and Entry by Nonenveloped Viruses. Trends Microbiol. 2018, 26 (6), 525– 537, DOI: 10.1016/j.tim.2017.09.010Google ScholarThere is no corresponding record for this reference.
- 18Mercer, J.; Schelhaas, M.; Helenius, A. Virus Entry by Endocytosis. Annu. Rev. Biochem. 2010, 79 (1), 803– 833, DOI: 10.1146/annurev-biochem-060208-104626Google Scholar18https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXpslSht7c%253D&md5=e083b9661ae42c15824b304f13c8e6fcVirus entry by endocytosisMercer, Jason; Schelhaas, Mario; Helenius, AriAnnual Review of Biochemistry (2010), 79 (), 803-833CODEN: ARBOAW; ISSN:0066-4154. (Annual Reviews Inc.)A review. Although viruses are simple in structure and compn., their interactions with host cells are complex. Merely to gain entry, animal viruses make use of a repertoire of cellular processes that involve hundreds of cellular proteins. Although some viruses have the capacity to penetrate into the cytosol directly through the plasma membrane, most depend on endocytic uptake, vesicular transport through the cytoplasm, and delivery to endosomes and other intracellular organelles. The internalization may involve clathrin-mediated endocytosis (CME), macropinocytosis, caveolar/lipid raft-mediated endocytosis, or a variety of other still poorly characterized mechanisms. This review focuses on the cell biol. of virus entry and the different strategies and endocytic mechanisms used by animal viruses.
- 19Pletan, M. L.; Tsai, B. Non-enveloped virus membrane penetration: New advances leading to new insights. PLoS Pathog. 2022, 18 (12), e1010948 DOI: 10.1371/journal.ppat.1010948Google ScholarThere is no corresponding record for this reference.
- 20Odegard, A.; Banerjee, M.; Johnson, J. E. Flock House Virus: A Model System for Understanding Non-Enveloped Virus Entry and Membrane Penetration. In Cell Entry by Non-Enveloped Viruses; Johnson, J. E., Ed.; Springer Berlin Heidelberg, 2010; pp 1– 22.Google ScholarThere is no corresponding record for this reference.
- 21Odegard, A. L.; Kwan, M. H.; Walukiewicz, H. E.; Banerjee, M.; Schneemann, A.; Johnson, J. E. Low endocytic pH and capsid protein autocleavage are critical components of Flock House virus cell entry. J. Virol. 2009, 83 (17), 8628– 8637, DOI: 10.1128/JVI.00873-09Google Scholar21https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhtFahurrJ&md5=ca9a32cff587b467ac1302d995eff723Low endocytic pH and capsid protein autocleavage are critical components of flock house virus cell entryOdegard, Amy L.; Kwan, Maggie H.; Walukiewicz, Hanna E.; Banerjee, Manidipa; Schneemann, Anette; Johnson, John E.Journal of Virology (2009), 83 (17), 8628-8637CODEN: JOVIAM; ISSN:0022-538X. (American Society for Microbiology)The process by which nonenveloped viruses cross cell membranes during host cell entry remains poorly defined; however, common themes are emerging. Here, we use correlated in vivo and in vitro studies to understand the mechanism of Flock House virus (FHV) entry and membrane penetration. We demonstrate that low endocytic pH is required for FHV infection, that exposure to acidic pH promotes FHV-mediated disruption of model membranes (liposomes), and particles exposed to low pH in vitro exhibit increased hydrophobicity. In addn., FHV particles perturbed by heating displayed a marked increase in liposome disruption, indicating that membrane-active regions of the capsid are exposed or released under these conditions. We also provide evidence that autoproteolytic cleavage, to generate the lipophilic γ peptide (4.4 kDa), is required for membrane penetration. Mutant, cleavage-defective particles failed to mediate liposome lysis, regardless of pH or heat treatment, suggesting that these particles are not able to expose or release the requisite membrane-active regions of the capsid, namely, the γ peptides. Based on these results, we propose an updated model for FHV entry in which (i) the virus enters the host cell by endocytosis, (ii) low pH within the endocytic pathway triggers the irreversible exposure or release of γ peptides from the virus particle, and (iii) the exposed/released γ peptides disrupt the endosomal membrane, facilitating translocation of viral RNA into the cytoplasm.
- 22Chen, N.-C.; Yoshimura, M.; Guan, H.-H.; Wang, T.-Y.; Misumi, Y.; Lin, C.-C.; Chuankhayan, P.; Nakagawa, A.; Chan, S. I.; Tsukihara, T. Crystal Structures of a Piscine Betanodavirus: Mechanisms of Capsid Assembly and Viral Infection. PLoS Pathog. 2015, 11 (10), e1005203, DOI: 10.1371/journal.ppat.1005203Google ScholarThere is no corresponding record for this reference.
- 23Gye, H. J.; Nishizawa, T. Sites responsible for infectivity and antigenicity on nervous necrosis virus (NNV) appear to be distinct. Sci. Rep. 2021, 11 (1), 3608 DOI: 10.1038/s41598-021-83078-3Google ScholarThere is no corresponding record for this reference.
- 24Gye, H. J.; Park, M.-J.; Kim, W.-S.; Oh, M.-J.; Nishizawa, T. Heat-denaturation of conformational structures on nervous necrosis virus for generating neutralization antibodies. Aquaculture 2018, 484, 65– 70, DOI: 10.1016/j.aquaculture.2017.10.034Google ScholarThere is no corresponding record for this reference.
- 25Nishizawa, T.; Mori, K.-i.; Furuhashi, M.; Nakai, T.; Furusawa, I.; Muroga, K. Comparison of the coat protein genes of five fish nodaviruses, the causative agents of viral nervous necrosis in marine fish. J. Gen. Virol. 1995, 76 (7), 1563– 1569, DOI: 10.1099/0022-1317-76-7-1563Google ScholarThere is no corresponding record for this reference.
- 26Iwamoto, T.; Okinaka, Y.; Mise, K.; Mori, K.-I.; Arimoto, M.; Okuno, T.; Nakai, T. Identification of Host-Specificity Determinants in Betanodaviruses by Using Reassortants between Striped Jack Nervous Necrosis Virus and Sevenband Grouper Nervous Necrosis Virus. J. Virol. 2004, 78 (3), 1256– 1262, DOI: 10.1128/JVI.78.3.1256-1262.2004Google ScholarThere is no corresponding record for this reference.
- 27Moreno, P.; Souto, S.; Leiva-Rebollo, R.; Borrego, J. J.; Bandín, I.; Alonso, M. C. Capsid amino acids at positions 247 and 270 are involved in the virulence of betanodaviruses to European sea bass. Sci. Rep. 2019, 9 (1), 14068 DOI: 10.1038/s41598-019-50622-1Google ScholarThere is no corresponding record for this reference.
- 28Nishizawa, T.; Lee, H. S.; Gye, H. J. Pocket structures of surface protrusions shared among serologically distinct nervous necrosis viruses (NNVs) were predicted in silico to bind to sialylated N-glycans, a host cellular receptor. Aquaculture 2023, 565, 739157, DOI: 10.1016/j.aquaculture.2022.739157Google ScholarThere is no corresponding record for this reference.
- 29Wang, C. H.; Chen, D. H.; Huang, S. H.; Wu, Y. M.; Chen, Y. Y.; Hwu, Y.; Bushnell, D.; Kornberg, R.; Chang, W. H. Sub-3 Å Cryo-EM Structures of Necrosis Virus Particles via the Use of Multipurpose TEM with Electron Counting Camera. Int. J. Mol. Sci. 2021, 22 (13), 6859, DOI: 10.3390/ijms22136859Google ScholarThere is no corresponding record for this reference.
- 30Chang, W. H.; Huang, S. H.; Lin, H. H.; Chung, S. C.; Tu, I. P. Cryo-EM Analyses Permit Visualization of Structural Polymorphism of Biological Macromolecules. Front. Bioinform. 2021, 1, 788308, DOI: 10.3389/fbinf.2021.788308Google ScholarThere is no corresponding record for this reference.
- 31Zhang, K.; Julius, D.; Cheng, Y. Structural snapshots of TRPV1 reveal mechanism of polymodal functionality. Cell 2021, 184 (20), 5138– 5150, DOI: 10.1016/j.cell.2021.08.012Google ScholarThere is no corresponding record for this reference.
- 32Chen, C.-Y.; Chang, Y.-C.; Lin, B.-L.; Huang, C.-H.; Tsai, M.-D. Temperature-Resolved Cryo-EM Uncovers Structural Bases of Temperature-Dependent Enzyme Functions. J. Am. Chem. Soc. 2019, 141 (51), 19983– 19987, DOI: 10.1021/jacs.9b10687Google Scholar32https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXitlygt7zO&md5=3866b861e15dbebe8e5657828643cbdaTemperature-Resolved Cryo-EM Uncovers Structural Bases of Temperature-Dependent Enzyme FunctionsChen, Chin-Yu; Chang, Yuan-Chih; Lin, Bo-Lin; Huang, Chun-Hsiang; Tsai, Ming-DawJournal of the American Chemical Society (2019), 141 (51), 19983-19987CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Protein functions are temp.-dependent, but protein structures are usually solved at a single (often low) temp. because of limitations on the conditions of crystal growth or protein vitrification. Here we demonstrate the feasibility of solving cryo-EM structures of proteins vitrified at high temps., solve 12 structures of an archaeal ketol-acid reductoisomerase (KARI) vitrified at 4-70 °C, and show that structures of both the Mg2+ form (KARI:2Mg2+) and its ternary complex (KARI:2Mg2+:NADH:inhibitor) are temp.-dependent in correlation with the temp. dependence of enzyme activity. Furthermore, structural analyses led to dissection of the induced-fit mechanism into ligand-induced and temp.-induced effects and to capture of temp.-resolved intermediates of the temp.-induced conformational change. The results also suggest that it is preferable to solve cryo-EM structures of protein complexes at functional temps. These studies should greatly expand the landscapes of protein structure-function relationships and enhance the mechanistic anal. of enzymic functions.
- 33Xie, J.; Li, K.; Gao, Y.; Huang, R.; Lai, Y.; Shi, Y.; Yang, S.; Zhu, G.; Zhang, Q.; He, J. Structural analysis and insertion study reveal the ideal sites for surface displaying foreign peptides on a betanodavirus-like particle. Vet. Res. 2016, 47 (1), 16, DOI: 10.1186/s13567-015-0294-9Google ScholarThere is no corresponding record for this reference.
- 34Štěrbová, P.; Wu, D.; Lou, Y.-C.; Wang, C.-H.; Chang, W.-H.; Tzou, D.-L. M. NMR assignments of protrusion domain of capsid protein from dragon grouper nervous necrosis virus. Biomol. NMR Assignments 2020, 14 (1), 63– 66, DOI: 10.1007/s12104-019-09921-xGoogle ScholarThere is no corresponding record for this reference.
- 35Walters, K. J.; Ferentz, A. E.; Hare, B. J.; Hidalgo, P.; Jasanoff, A.; Matsuo, H.; Wagner, G. Characterizing protein-protein complexes and oligomers by nuclear magnetic resonance spectroscopy. In Methods in Enzymology; Elsevier, 2001; Vol. 339, pp 238– 258. DOI: 10.1016/S0076-6879(01)39316-3 .Google ScholarThere is no corresponding record for this reference.
- 36Williamson, M. P. Using chemical shift perturbation to characterise ligand binding. Prog. Nucl. Magn. Reson. Spectrosc. 2013, 73, 1– 16, DOI: 10.1016/j.pnmrs.2013.02.001Google Scholar36https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXht1yrsbzE&md5=74fc06bb4f4b3df369619bbf730016ceUsing chemical shift perturbation to characterise ligand bindingWilliamson, Mike P.Progress in Nuclear Magnetic Resonance Spectroscopy (2013), 73 (), 1-16CODEN: PNMRAT; ISSN:0079-6565. (Elsevier B.V.)A review. Chem. shift perturbation (CSP, chem. shift mapping or complexation-induced changes in chem. shift, CIS) follows changes in the chem. shifts of a protein when a ligand is added, and uses these to det. the location of the binding site, the affinity of the ligand, and/or possibly the structure of the complex. A key factor in detg. the appearance of spectra during a titrn. is the exchange rate between free and bound, or more specifically the off-rate koff. When koff is greater than the chem. shift difference between free and bound, which typically equates to an affinity Kd weaker than about 3 μM, then exchange is fast on the chem. shift timescale. Under these circumstances, the obsd. shift is the population-weighted av. of free and bound, which allows Kd to be detd. from measurement of peak positions, provided the measurements are made appropriately. 1H shifts are influenced to a large extent by through-space interactions, whereas 13Cα and 13Cβ shifts are influenced more by through-bond effects. 15N and 13C' shifts are influenced both by through-bond and by through-space (hydrogen bonding) interactions. For detg. the location of a bound ligand on the basis of shift change, the most appropriate method is therefore usually to measure 15N HSQC spectra, calc. the geometrical distance moved by the peak, weighting 15N shifts by a factor of about 0.14 compared to 1H shifts, and select those residues for which the weighted shift change is larger than the std. deviation of the shift for all residues. Other methods are discussed, in particular the measurement of 13CH3 signals. Slow to intermediate exchange rates lead to line broadening, and make Kd values very difficult to obtain. There is no good way to distinguish changes in chem. shift due to direct binding of the ligand from changes in chem. shift due to allosteric change. Ligand binding at multiple sites can often be characterized, by simultaneous fitting of many measured shift changes, or more simply by adding substoichiometric amts. of ligand. The chem. shift changes can be used as restraints for docking ligand onto protein. By use of quant. calcns. of ligand-induced chem. shift changes, it is becoming possible to det. not just the position but also the orientation of ligands.
- 37Krissinel, E.; Henrick, K. Inference of Macromolecular Assemblies from Crystalline State. J. Mol. Biol. 2007, 372 (3), 774– 797, DOI: 10.1016/j.jmb.2007.05.022Google Scholar37https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXpvFGktb8%253D&md5=a5c764cfc7dc129f53ddc31ef9d475faInference of Macromolecular Assemblies from Crystalline StateKrissinel, Evgeny; Henrick, KimJournal of Molecular Biology (2007), 372 (3), 774-797CODEN: JMOBAK; ISSN:0022-2836. (Elsevier Ltd.)The authors discuss basic phys.-chem. principles underlying the formation of stable macromol. complexes, which in many cases are likely to be the biol. units performing a certain physiol. function. The authors also consider available theor. approaches to the calcn. of macromol. affinity and entropy of complexation. The latter is shown to play an important role and make a major effect on complex size and symmetry. The authors develop a new method, based on chem. thermodn., for automatic detection of macromol. assemblies in the Protein Data Bank (PDB) entries that are the results of x-ray diffraction expts. As found, biol. units may be recovered at 80-90% success rate, which makes x-ray crystallog. an important source of exptl. data on macromol. complexes and protein-protein interactions. The method is implemented as a public WWW service (http://www.ebi.ac.uk/msd-srv/prot_int/pistart.html).
- 38Kampmann, T.; Mueller, D. S.; Mark, A. E.; Young, P. R.; Kobe, B. The Role of histidine residues in low-pH-mediated viral membrane fusion. Structure 2006, 14 (10), 1481– 1487, DOI: 10.1016/j.str.2006.07.011Google Scholar38https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28XhtVCrtrnN&md5=ffb8d25828217e82747ef2838f53f028The Role of Histidine Residues in Low-pH-Mediated Viral Membrane FusionKampmann, Thorsten; Mueller, Daniela S.; Mark, Alan E.; Young, Paul R.; Kobe, BostjanStructure (Cambridge, MA, United States) (2006), 14 (10), 1481-1487CODEN: STRUE6; ISSN:0969-2126. (Cell Press)A central event in the invasion of a host cell by an enveloped virus is the fusion of viral and cell membranes. For many viruses, membrane fusion is driven by specific viral surface proteins that undergo large-scale conformational rearrangements, triggered by exposure to low pH in the endosome upon internalization. Here, we present evidence suggesting that in both class I (helical hairpin proteins) and class II (β-structure-rich proteins) pH-dependent fusion proteins the protonation of specific histidine residues triggers fusion via an analogous mol. mechanism. These histidines are located in the vicinity of pos. charged residues in the prefusion conformation, and they subsequently form salt bridges with neg. charged residues in the postfusion conformation. The mol. surfaces involved in the corresponding structural rearrangements leading to fusion are highly conserved and thus might provide a suitable common target for the design of antivirals, which could be active against a diverse range of pathogenic viruses.
- 39The PyMOL Molecular Graphics System, Version 1.8, Schrodinger, LLC, 2015.Google ScholarThere is no corresponding record for this reference.
- 40Wedemeyer, W. J.; Welker, E.; Scheraga, H. A. Proline cis-trans isomerization and protein folding. Biochemistry 2002, 41 (50), 14637– 14644, DOI: 10.1021/bi020574bGoogle Scholar40https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD38XotlOmsLw%253D&md5=f05d94502166b940f92d215d49352b6fProline Cis-Trans Isomerization and Protein FoldingWedemeyer, William J.; Welker, Ervin; Scheraga, Harold A.Biochemistry (2002), 41 (50), 14637-14644CODEN: BICHAW; ISSN:0006-2960. (American Chemical Society)A review. Proline cis-trans isomerization plays a key role in the rate-detg. steps of protein folding. The energetic origin of this isomerization process is summarized, and the folding and unfolding of disulfide-intact bovine pancreatic RNase A is used as an example to illustrate the kinetics and structural features of conformational changes from the heterogeneous unfolded state (consisting of cis and trans isomers of X-Pro peptide groups) to the native structure in which only one set of proline isomers is present.
- 41Shen, Y.; Bax, A. Prediction of Xaa-Pro peptide bond conformation from sequence and chemical shifts. J. Biomol. NMR 2010, 46 (3), 199– 204, DOI: 10.1007/s10858-009-9395-yGoogle Scholar41https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXivV2ku70%253D&md5=30d471955095ff1da608e18b1d3321d4Prediction of Xaa-Pro peptide bond conformation from sequence and chemical shiftsShen, Yang; Bax, AdJournal of Biomolecular NMR (2010), 46 (3), 199-204CODEN: JBNME9; ISSN:0925-2738. (Springer)The authors present a program, named Promega, to predict the Xaa-Pro peptide bond conformation on the basis of backbone chem. shifts and the amino acid sequence. Using a chem. shift database of proteins of known structure together with the PDB-extd. amino acid preference of cis Xaa-Pro peptide bonds, a cis/trans probability score is calcd. from the backbone and 13Cβ chem. shifts of the proline and its neighboring residues. For an arbitrary no. of input chem. shifts, which may include Pro-13Cγ, Promega calcs. the statistical probability that a Xaa-Pro peptide bond is cis. Besides its potential as a validation tool, Promega is particularly useful for studies of larger proteins where Pro-13Cγ assignments can be challenging, and for on-going efforts to det. protein structures exclusively on the basis of backbone and 13Cβ chem. shifts.
- 42Gye, H. J.; Nishizawa, T. Analysis of sialylated N-linked glycans on fish cell lines permissive to nervous necrosis virus for predicting cellular receptors of the virus. Aquaculture 2022, 555, 738198 DOI: 10.1016/j.aquaculture.2022.738198Google ScholarThere is no corresponding record for this reference.
- 43Rodríguez, Y.; Cardoze, S. M.; Obineche, O. W.; Melo, C.; Persaud, A.; Fernández Romero, J. A. Small Molecules Targeting SARS-CoV-2 Spike Glycoprotein Receptor-Binding Domain. ACS Omega 2022, 7 (33), 28779– 28789, DOI: 10.1021/acsomega.2c00844Google Scholar43https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB38XitVKktrbO&md5=4c573b63f6c38170fbe0bff8533dd57bSmall Molecules Targeting SARS-CoV-2 Spike Glycoprotein Receptor-Binding DomainRodriguez, Yoel; Cardoze, Scarlet Martinez; Obineche, Onyinyechi W.; Melo, Claudia; Persaud, Ashanna; Fernandez Romero, Jose A.ACS Omega (2022), 7 (33), 28779-28789CODEN: ACSODF; ISSN:2470-1343. (American Chemical Society)The severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) has caused the coronavirus disease 2019 (COVID-19) pandemic. Several variants of SARS-CoV-2 have emerged worldwide. These variants show different transmissibility infectivity due to mutations in the viral spike (S) glycoprotein that interacts with the human angiotensin-converting enzyme 2 (hACE2) receptor and facilitates viral entry into target cells. Despite the effective SARS-CoV-2 vaccines, we still need to identify selective antivirals, and the S glycoprotein is a key target to neutralize the virus. We hypothesize that small mols. could disrupt the interaction of S glycoprotein with hACE2 and inhibit viral entry. We analyzed the S glycoprotein-hACE2 complex structure (PDB: 7DF4) and created models for different viral variants using visual mol. dynamics (VMD) and mol. operating environment (MOE) programs. Moreover, we started the hits search by performing structure-based mol. docking virtual screening of com. available small mols. against S glycoprotein models using OEDocking FRED-4.0.0.0 software. The FRED-4.0.0.0 Chemguass4 scoring function was used to rank the small mols. based on their affinities. The best candidate compds. were purchased and tested using a std. SARS-CoV-2 pseudotyped cell-based bioassay to investigate their antiviral activity. Three of these compds., alone or in combination, showed antiviral selectivity. These small mols. may lead to an effective antiviral treatment or serve as probes to better understand the biol. of SARS-CoV-2.
- 44Chang, J.-S.; Chi, S.-C. GHSC70 is involved in the cellular entry of nervous necrosis virus. J. Virol. 2015, 89 (1), 61– 70, DOI: 10.1128/JVI.02523-14Google ScholarThere is no corresponding record for this reference.
- 45Zhang, W.; Jia, K.; Jia, P.; Xiang, Y.; Lu, X.; Liu, W.; Yi, M. Marine medaka heat shock protein 90ab1 is a receptor for red-spotted grouper nervous necrosis virus and promotes virus internalization through clathrin-mediated endocytosis. PLoS Pathog. 2020, 16 (7), e1008668 DOI: 10.1371/journal.ppat.1008668Google ScholarThere is no corresponding record for this reference.
- 46Krishnan, R.; Qadiri, S. S. N.; Oh, M.-J. Functional characterization of seven-band grouper immunoglobulin like cell adhesion molecule, Nectin4 as a cellular receptor for nervous necrosis virus. Fish Shellfish Immunol. 2019, 93, 720– 725, DOI: 10.1016/j.fsi.2019.08.019Google ScholarThere is no corresponding record for this reference.
- 47Ito, Y.; Okinaka, Y.; Mori, K. I.; Sugaya, T.; Nishioka, T.; Oka, M.; Nakai, T. Variable region of betanodavirus RNA2 is sufficient to determine host specificity. Dis. Aquat. Org. 2008, 79 (3), 199– 205, DOI: 10.3354/dao01906Google ScholarThere is no corresponding record for this reference.
- 48Staring, J.; Raaben, M.; Brummelkamp, T. R. Viral escape from endosomes and host detection at a glance. J. Cell Sci. 2018, 131 (15), jcs216259 DOI: 10.1242/jcs.216259Google ScholarThere is no corresponding record for this reference.
- 49Song, C.; Takai-Todaka, R.; Miki, M.; Haga, K.; Fujimoto, A.; Ishiyama, R.; Oikawa, K.; Yokoyama, M.; Miyazaki, N.; Iwasaki, K. Dynamic rotation of the protruding domain enhances the infectivity of norovirus. PLoS Pathog. 2020, 16 (7), e1008619 DOI: 10.1371/journal.ppat.1008619Google ScholarThere is no corresponding record for this reference.
- 50Hu, L.; Salmen, W.; Chen, R.; Zhou, Y.; Neill, F.; Crowe, J. E.; Atmar, R. L.; Estes, M. K.; Prasad, B. V. V. Atomic structure of the predominant GII.4 human norovirus capsid reveals novel stability and plasticity. Nat. Commun. 2022, 13 (1), 1241 DOI: 10.1038/s41467-022-28757-zGoogle ScholarThere is no corresponding record for this reference.
- 51Goto, Y.; Takahashi, N.; Fink, A. L. (1990). Mechanism of acid-induced folding of proteins. Biochemistry 1990, 29 (14), 3480– 3488, DOI: 10.1021/bi00466a009Google Scholar51https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3cXhsFeqs7w%253D&md5=48e21a0012913f7c73f4748b2c62a142Mechanism of acid-induced folding of proteinsGoto, Yuji; Takahashi, Nobuaki; Fink, Anthony L.Biochemistry (1990), 29 (14), 3480-8CODEN: BICHAW; ISSN:0006-2960.It was previously shown that β-lactamase, cytochrome c (I), and apomyoglobin (II) are maximally unfolded at pH 2 under conditions of low ionic strength, but that a further decrease in pH, by increasing the concn. of HCl, refolds the proteins to the A state with properties similar to those of a molten globule state. To understand the mechanism of acid-induced refolding of protein structure, the effects of various strong acids and their neutral salts on the acid-unfolded states of ferri-I and II were studied. The conformational transition of I was monitored at 20° by using changes in the far-UV CD and in the Soret absorption at 394 nm, and that of II was monitored by changes in the far-UV CD. Various strong acids (i.e., H2SO4, HClO4, HNO3, TCA, and trifluoroacetic acid) refolded acid-unfolded I and II to the A states as was the case with HCl. For both proteins, neutral salts of these acids caused similar conformational transitions, confirming that the anions are responsible for bringing about the transition. The order of effectiveness of anions was ferricyanide > ferrocyanide > SO42- > TCA- > SCN- > ClO4- > I- > NO3- > trifluoroacetate > Br- > Cl-. This series was similar to the electroselectivity series of anions toward the anion-exchange resins, showing that preferential binding of anions to the A states causes the conformational transitions.
- 52Güthe, S.; Kapinos, L.; Möglich, A.; Meier, S.; Grzesiek, S.; Kiefhaber, T. Very fast folding and association of a trimerization domain from bacteriophage T4 fibritin. J. Mol. Biol. 2004, 337 (4), 905– 915, DOI: 10.1016/j.jmb.2004.02.020Google ScholarThere is no corresponding record for this reference.
- 53So, M.; Hall, D.; Goto, Y. Revisiting supersaturation as a factor determining amyloid fibrillation. Curr. Opin. Struct. Biol. 2016, 36, 32– 39, DOI: 10.1016/j.sbi.2015.11.009Google Scholar53https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXitVSkurjP&md5=01c17457e922fbefc12715f0bf000e1dRevisiting supersaturation as a factor determining amyloid fibrillationSo, Masatomo; Hall, Damien; Goto, YujiCurrent Opinion in Structural Biology (2016), 36 (), 32-39CODEN: COSBEF; ISSN:0959-440X. (Elsevier Ltd.)Amyloid fibrils involved in various diseases are formed by a nucleation-growth mechanism, similar to the crystn. of solutes from soln. Soly. and supersatn. are two of the most important factors detg. crystn. of solutes. Moreover, crystn. competes with glass formation in which solutes collapse into amorphous aggregates. Recent studies on the formation of amyloid fibrils and amorphous aggregates indicate that the partition between distinct types of aggregates can be rationally explained by a kinetic and thermodn. competition between them. Understanding the role of supersatn. in detg. aggregation-based phase transitions of denatured proteins provides an important complementary point of view to structural studies of protein aggregates.
- 54Shih, T. C.; Ho, L. P.; Chou, H. Y.; Wu, J. L.; Pai, T. W. Comprehensive Linear Epitope Prediction System for Host Specificity in Nodaviridae. Viruses 2022, 14 (7), 1357, DOI: 10.3390/v14071357Google ScholarThere is no corresponding record for this reference.
- 55Zhang, Z.; Xing, J.; Tang, X.; Sheng, X.; Chi, H.; Zhan, W. Development and characterization of monoclonal antibodies against red-spotted grouper nervous necrosis virus and their neutralizing potency in vitro. Aquaculture 2022, 560, 738562 DOI: 10.1016/j.aquaculture.2022.738562Google ScholarThere is no corresponding record for this reference.
- 56Huang, S.; Wu, Y.; Su, L.; Su, T.; Zhou, Q.; Zhang, J.; Zhao, Z.; Weng, S.; He, J.; Xie, J. A single-chain variable fragment antibody exerts anti-nervous necrosis virus activity by irreversible binding. Aquaculture 2022, 552, 738001 DOI: 10.1016/j.aquaculture.2022.738001Google ScholarThere is no corresponding record for this reference.
- 57Graham, B. S.; Gilman, M. S. A.; McLellan, J. S. Structure-Based Vaccine Antigen Design. Annu. Rev. Med. 2019, 70, 91– 104, DOI: 10.1146/annurev-med-121217-094234Google Scholar57https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXhvF2ltrw%253D&md5=ccbef186198ab5f51fe6aa228b440734Structure-Based Vaccine Antigen DesignGraham, Barney S.; Gilman, Morgan S. A.; McLellan, Jason S.Annual Review of Medicine (2019), 70 (), 91-104CODEN: ARMCAH; ISSN:0066-4219. (Annual Reviews)A review. Enabled by new approaches for rapid identification and selection of human monoclonal antibodies, at.-level structural information for viral surface proteins, and capacity for precision engineering of protein immunogens and self-assembling nanoparticles, a new era of antigen design and display options has evolved. While HIV-1 vaccine development has been a driving force behind these technologies and concepts, clin. proof-of-concept for structure-based vaccine design may first be achieved for respiratory syncytial virus (RSV), where conformation-dependent access to neutralization-sensitive epitopes on the fusion glycoprotein dets. the capacity to induce potent neutralizing activity. Success with RSV has motivated structure-based stabilization of other class I viral fusion proteins for use as immunogens and demonstrated the importance of structural information for developing vaccines against other viral pathogens, particularly difficult targets that have resisted prior vaccine development efforts. Solving viral surface protein structures also supports rapid vaccine antigen design and application of platform manufg. approaches for emerging pathogens.
- 58Liang, J.; Edelsbrunner, H.; Woodward, C. Anatomy of protein pockets and cavities: measurement of binding site geometry and implications for ligand design. Protein Sci. 1998, 7 (9), 1884– 1897, DOI: 10.1002/pro.5560070905Google Scholar58https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1cXmtFGjsLo%253D&md5=a6f62e03de03da0a6e51361de7a08765Anatomy of protein pockets and cavities: measurement of binding site geometry and implications for ligand designLiang, Jie; Edelsbrunner, Herbert; Woodward, ClareProtein Science (1998), 7 (9), 1884-1897CODEN: PRCIEI; ISSN:0961-8368. (Cambridge University Press)Identification and size characterization of surface pockets and occluded cavities are initial steps in protein structure-based ligand design. A new program, CAST, for automatically locating and measuring protein pockets and cavities, is based on precise computational geometry methods, including alpha shape and discrete flow theory. CAST identifies and measures pockets and pocket mouth openings, as well as cavities. The program specifies the atoms lining pockets, pocket openings, and buried cavities; the vol. and area of pockets and cavities; and the area and circumference of mouth openings. CAST anal. of over 100 proteins has been carried out; proteins examd. include a set of 51 monomeric enzyme-ligand structures, several elastase-inhibitor complexes, the FK506 binding protein, 30 HIV-1 protease-inhibitor complexes, and a no. of small and large protein inhibitors. Medium-sized globular proteins typically have 10-20 pockets/cavities. Most often, binding sites are pockets with 1-2 mouth openings; much less frequently they are cavities. Ligand binding pockets vary widely in size, most within the range 102-103 Å3. Statistical anal. reveals that the no. of pockets and cavities is correlated with protein size, but there is no correlation between the size of the protein and the size of binding sites. Most frequently, the largest pocket/cavity is the active site, but there are a no. of instructive exceptions. Ligand vol. and binding site vol. are somewhat correlated when binding site vol. is ≤700 Å3, but the ligand seldom occupies the entire site. Auxiliary pockets near the active site have been suggested as addnl. binding surface for designed ligands (Mattos C et al., 1994, Nat Struct Biol 1:55-58). Anal. of elastase-inhibitor complexes suggests that CAST can identify ancillary pockets suitable for recruitment in ligand design strategies. Anal. of the FK506 binding protein, and of compds. developed in SAR by NMR (Shuker SB et al., 1996, Science 274:1531-1534), indicates that CAST pocket computation may provide a priori identification of target proteins for linked-fragment design. CAST anal. of 30 HIV-1 protease-inhibitor complexes shows that the flexible active site pocket can vary over a range of 853-1,566 Å3, and that there are two pockets near or adjoining the active site that may be recruited for ligand design.
- 59Naceri, S.; Marc, D.; Blot, R.; Flatters, D.; Camproux, A. C. Druggable Pockets at the RNA Interface Region of Influenza A Virus NS1 Protein Are Conserved across Sequence Variants from Distinct Subtypes. Biomolecules 2023, 13 (1), 64, DOI: 10.3390/biom13010064Google ScholarThere is no corresponding record for this reference.
- 60Gavenonis, J.; Sheneman, B. A.; Siegert, T. R.; Eshelman, M. R.; Kritzer, J. A. Comprehensive analysis of loops at protein-protein interfaces for macrocycle design. Nat. Chem. Biol. 2014, 10 (9), 716– 722, DOI: 10.1038/nchembio.1580Google Scholar60https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhtFygs7jE&md5=bf07bd724f173ea74467ab7cd5aedb6aComprehensive analysis of loops at protein-protein interfaces for macrocycle designGavenonis, Jason; Sheneman, Bradley A.; Siegert, Timothy R.; Eshelman, Matthew R.; Kritzer, Joshua A.Nature Chemical Biology (2014), 10 (9), 716-722CODEN: NCBABT; ISSN:1552-4450. (Nature Publishing Group)Inhibiting protein-protein interactions (PPIs) with synthetic mols. remains a frontier of chem. biol. Many PPIs have been successfully targeted by mimicking α-helixes at interfaces, but most PPIs are mediated by nonhelical, nonstrand peptide loops. We sought to comprehensively identify and analyze these loop-mediated PPIs by writing and implementing LoopFinder, a customizable program that can identify loop-mediated PPIs within all of the protein-protein complexes in the Protein Data Bank. Comprehensive anal. of the entire set of 25,005 interface loops revealed common structural motifs and unique features that distinguish loop-mediated PPIs from other PPIs. 'Hot loops', named in analogy to protein hot spots, were identified as loops with favorable properties for mimicry using synthetic mols. The hot loops and their binding partners represent new and promising PPIs for the development of macrocycle and constrained peptide inhibitors.
- 61Corbi-Verge, C.; Kim, P. M. Motif mediated protein-protein interactions as drug targets. Cell Commun. Signal 2016, 14, 8 DOI: 10.1186/s12964-016-0131-4Google ScholarThere is no corresponding record for this reference.
- 62Marković, V.; Szczepańska, A.; Berlicki, Ł. Antiviral Protein-Protein Interaction Inhibitors. J. Med. Chem. 2024, 67 (5), 3205– 3231, DOI: 10.1021/acs.jmedchem.3c01543Google ScholarThere is no corresponding record for this reference.
- 63Zheng, 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 (4), 331– 332, DOI: 10.1038/nmeth.4193Google Scholar63https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXjt1ags7g%253D&md5=5f4e225ef8123dacd8475d526175e1d2MotionCor2: anisotropic correction of beam-induced motion for improved cryo-electron microscopyZheng, Shawn Q.; Palovcak, Eugene; Armache, Jean-Paul; Verba, Kliment A.; Cheng, Yifan; Agard, David A.Nature Methods (2017), 14 (4), 331-332CODEN: NMAEA3; ISSN:1548-7091. (Nature Publishing Group)A review on anisotropic correction of beam-induced motion for improved cryo-electron microscopy. Here we describe MotionCor2, a software tool for anisotropic correction of beam-induced motion. Overall, MotionCor2 is extremely robust and sufficiently accurate at correcting local motions so that the very time-consuming and computationally intensive particle polishing in RELION can be skipped, importantly, it also works on a wide range of data sets, including cryo tomog. tilt series.
- 64Rohou, A.; Grigorieff, N. CTFFIND4: Fast and accurate defocus estimation from electron micrographs. J. Struct. Biol. 2015, 192 (2), 216– 221, DOI: 10.1016/j.jsb.2015.08.008Google Scholar64https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC287js1Whsg%253D%253D&md5=8500953ad4898ae82de6f8cdc95832cfCTFFIND4: Fast and accurate defocus estimation from electron micrographsRohou Alexis; Grigorieff NikolausJournal of structural biology (2015), 192 (2), 216-21 ISSN:.CTFFIND is a widely-used program for the estimation of objective lens defocus parameters from transmission electron micrographs. Defocus parameters are estimated by fitting a model of the microscope's contrast transfer function (CTF) to an image's amplitude spectrum. Here we describe modifications to the algorithm which make it significantly faster and more suitable for use with images collected using modern technologies such as dose fractionation and phase plates. We show that this new version preserves the accuracy of the original algorithm while allowing for higher throughput. We also describe a measure of the quality of the fit as a function of spatial frequency and suggest this can be used to define the highest resolution at which CTF oscillations were successfully modeled.
- 65Punjani, A.; Rubinstein, J. L.; Fleet, D. J.; Brubaker, M. A. cryoSPARC: algorithms for rapid unsupervised cryo-EM structure determination. Nat. Methods 2017, 14 (3), 290– 296, DOI: 10.1038/nmeth.4169Google Scholar65https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXitlGisbs%253D&md5=95d468147707707e70ac0ad38dd6ebf6cryoSPARC: algorithms for rapid unsupervised cryo-EM structure determinationPunjani, Ali; Rubinstein, John L.; Fleet, David J.; Brubaker, Marcus A.Nature Methods (2017), 14 (3), 290-296CODEN: NMAEA3; ISSN:1548-7091. (Nature Publishing Group)Single-particle electron cryomicroscopy (cryo-EM) is a powerful method for detg. the structures of biol. macromols. With automated microscopes, cryo-EM data can often be obtained in a few days. However, processing cryo-EM image data to reveal heterogeneity in the protein structure and to refine 3D maps to high resoln. frequently becomes a severe bottleneck, requiring expert intervention, prior structural knowledge, and weeks of calcns. on expensive computer clusters. Here we show that stochastic gradient descent (SGD) and branch-and-bound max. likelihood optimization algorithms permit the major steps in cryo-EM structure detn. to be performed in hours or minutes on an inexpensive desktop computer. Furthermore, SGD with Bayesian marginalization allows ab initio 3D classification, enabling automated anal. and discovery of unexpected structures without bias from a ref. map. These algorithms are combined in a user-friendly computer program named cryoSPARC (http://www.cryosparc.com).
- 66Yang, Z.; Lasker, K.; Schneidman-Duhovny, D.; Webb, B.; Huang, C. C.; Pettersen, E. F.; Goddard, T. D.; Meng, E. C.; Sali, A.; Ferrin, T. E. UCSF Chimera, MODELLER, and IMP: an integrated modeling system. J. Struct Biol. 2012, 179 (3), 269– 278, DOI: 10.1016/j.jsb.2011.09.006Google ScholarThere is no corresponding record for this reference.
- 67Dam, J.; Schuck, P. Calculating Sedimentation Coefficient Distributions by Direct Modeling of Sedimentation Velocity Concentration Profiles. In Methods in Enzymology; Academic Press, 2004; Vol. 384, pp 185– 212.Google ScholarThere is no corresponding record for this reference.
- 68Ribaric, S.; Peterec, D.; Sketelj, J. Computer aided data acquisition and analysis of acetlycholinesterase velocity sedimentation profiles. Comput. Methods Program. Biomed. 1996, 49 (2), 149– 156, DOI: 10.1016/0169-2607(96)01719-1Google ScholarThere is no corresponding record for this reference.
- 69Johnson, B. A.; Blevins, R. A. NMR View: A computer program for the visualization and analysis of NMR data. J. Biomol. NMR 1994, 4 (5), 603– 614, DOI: 10.1007/BF00404272Google Scholar69https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2cXmt1Gkurw%253D&md5=6a93a0d9704fbbb654abc192203c68aeNMRView: a computer program for the visualization and analysis of NMR dataJohnson, Bruce A.; Blevins, Richard A.Journal of Biomolecular NMR (1994), 4 (5), 603-14CODEN: JBNME9; ISSN:0925-2738.NMRView is a computer program designed for the visualization and anal. of NMR data. It allows the user to interact with a practically unlimited no. of 2D, 3D and 4D NMR data files. Any no. of spectral windows can be displayed on the screen in any size and location. Automatic peak picking and facilitated peak anal. features are included to aid in the assignment of complex NMR spectra. NMR View provides structure anal. features and data transfer to and from structural generation programs, allowing for a tight coupling between the spectral anal. and structural generation.,. Visual correlation between structures and spectra can be done with the Mol. Data Viewer, a mol. graphics program with bidirectional communication to NMR View. The used interface can be customized and a command language is provided to allow for the automation of various tasks.
- 70Schwieters, C. D.; Kuszewski, J. J.; Marius Clore, G. Using Xplor–NIH for NMR molecular structure determination. Prog. Nucl. Magn. Reson. Spectrosc. 2006, 48 (1), 47– 62, DOI: 10.1016/j.pnmrs.2005.10.001Google Scholar70https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28XjtVOiurY%253D&md5=56257dc51ddafdf8c467d30aef191ee2Using Xplor-NIH for NMR molecular structure determinationSchwieters, Charles D.; Kuszewski, John J.; Clore, G. MariusProgress in Nuclear Magnetic Resonance Spectroscopy (2006), 48 (1), 47-62CODEN: PNMRAT; ISSN:0079-6565. (Elsevier B.V.)A review of the title program for computerized structure detn.
- 71Maciejewski, M. W.; Schuyler, A. D.; Gryk, M. R.; Moraru, I. I.; Romero, P. R.; Ulrich, E. L.; Eghbalnia, H. R.; Livny, M.; Delaglio, F.; Hoch, J. C. NMRbox: A Resource for Biomolecular NMR Computation. Biophys. J. 2017, 112 (8), 1529– 1534, DOI: 10.1016/j.bpj.2017.03.011Google Scholar71https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXlvFWgtL8%253D&md5=a990d67aed35b62cd010f0f5396d0f82NMRbox: A Resource for Biomolecular NMR ComputationMaciejewski, Mark W.; Schuyler, Adam D.; Gryk, Michael R.; Moraru, Ion I.; Romero, Pedro R.; Ulrich, Eldon L.; Eghbalnia, Hamid R.; Livny, Miron; Delaglio, Frank; Hoch, Jeffrey C.Biophysical Journal (2017), 112 (8), 1529-1534CODEN: BIOJAU; ISSN:0006-3495. (Cell Press)Advances in computation have been enabling many recent advances in biomol. applications of NMR. Due to the wide diversity of applications of NMR, the no. and variety of software packages for processing and analyzing NMR data is quite large, with labs relying on dozens, if not hundreds of software packages. Discovery, acquisition, installation, and maintenance of all these packages is a burdensome task. Because the majority of software packages originate in academic labs, persistence of the software is compromised when developers graduate, funding ceases, or investigators turn to other projects. To simplify access to and use of biomol. NMR software, foster persistence, and enhance reproducibility of computational workflows, the authors have developed NMRbox, a shared resource for NMR software and computation. NMRbox employs virtualization to provide a comprehensive software environment preconfigured with hundreds of software packages, available as a downloadable virtual machine or as a Platform-as-a-Service supported by a dedicated compute cloud. Ongoing development includes a metadata harvester to regularize, annotate, and preserve workflows and facilitate and enhance data depositions to BioMagResBank, and tools for Bayesian inference to enhance the robustness and extensibility of computational analyses. In addn. to facilitating use and preservation of the rich and dynamic software environment for biomol. NMR, NMRbox fosters the development and deployment of a new class of metasoftware packages. NMRbox is freely available to not-for-profit users.
- 72Shen, Y.; Delaglio, F.; Cornilescu, G.; Bax, A. TALOS+: a hybrid method for predicting protein backbone torsion angles from NMR chemical shifts. J. Biomol. NMR 2009, 44 (4), 213– 223, DOI: 10.1007/s10858-009-9333-zGoogle Scholar72https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXovFajtr8%253D&md5=ee3f8647a02e19937988056fd7af04b2TALOS+: a hybrid method for predicting protein backbone torsion angles from NMR chemical shiftsShen, Yang; Delaglio, Frank; Cornilescu, Gabriel; Bax, AdJournal of Biomolecular NMR (2009), 44 (4), 213-223CODEN: JBNME9; ISSN:0925-2738. (Springer)NMR chem. shifts in proteins depend strongly on local structure. The program TALOS establishes an empirical relation between 13C, 15N and 1H chem. shifts and backbone torsion angles .vphi. and ψ. Extension of the original 20-protein database to 200 proteins increased the fraction of residues for which backbone angles could be predicted from 65 to 74%, while reducing the error rate from 3 to 2.5%. Addn. of a two-layer neural network filter to the database fragment selection process forms the basis for a new program, TALOS+, which further enhances the prediction rate to 88.5%, without increasing the error rate. Excluding the 2.5% of residues for which TALOS+ makes predictions that strongly differ from those obsd. in the cryst. state, the accuracy of predicted .vphi. and ψ angles, equals ±13°. Large discrepancies between predictions and crystal structures are primarily limited to loop regions, and for the few cases where multiple X-ray structures are available such residues are often found in different states in the different structures. The TALOS+ output includes predictions for individual residues with missing chem. shifts, and the neural network component of the program also predicts secondary structure with good accuracy.
- 73Tian, Y.; Schwieters, C. D.; Opella, S. J.; Marassi, F. M. A practical implicit solvent potential for NMR structure calculation. J. Magn. Reson. 2014, 243, 54– 64, DOI: 10.1016/j.jmr.2014.03.011Google Scholar73https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXovVahu70%253D&md5=68ad01c9ceb6078a4291b0d6e96408e8A practical implicit solvent potential for NMR structure calculationTian, Ye; Schwieters, Charles D.; Opella, Stanley J.; Marassi, Francesca M.Journal of Magnetic Resonance (2014), 243 (), 54-64CODEN: JMARF3; ISSN:1090-7807. (Elsevier B.V.)The benefits of protein structure refinement in water are well documented. However, performing structure refinement with explicit at. representation of the solvent mols. is computationally expensive and impractical for NMR-restrained structure calcns. that start from completely extended polypeptide templates. Here we describe a new implicit solvation potential, EEFx (Effective Energy Function for XPLOR-NIH), for NMR-restrained structure calcns. of proteins in XPLOR-NIH. The key components of EEFx are an energy term for solvation energy that works together with other nonbonded energy functions, and a dedicated force field for conformational and nonbonded protein interaction parameters. The initial results obtained with EEFx show that significant improvements in structural quality can be obtained. EEFx is computationally efficient and can be used both to fold and refine structures. Overall, EEFx improves the quality of protein conformation and nonbonded at. interactions. Moreover, such benefits are accompanied by enhanced structural precision and enhanced structural accuracy, reflected in improved agreement with the cross-validated dipolar coupling data. Finally, implementation of EEFx calcns. is straightforward and computationally efficient. Overall, EEFx provides a useful method for the practical calcn. of exptl. protein structures in a phys. realistic environment.
- 74Gore, S.; Sanz García, E.; Hendrickx, P. M. S.; Gutmanas, A.; Westbrook, J. D.; Yang, H.; Feng, Z.; Baskaran, K.; Berrisford, J. M.; Hudson, B. P. Validation of Structures in the Protein Data Bank. Structure 2017, 25 (12), 1916– 1927, DOI: 10.1016/j.str.2017.10.009Google Scholar74https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXhvVOitbnE&md5=680c220ca1354e8a089638ae978501d3Validation of Structures in the Protein Data BankGore, Swanand; Sanz Garcia, Eduardo; Hendrickx, Pieter M. S.; Gutmanas, Aleksandras; Westbrook, John D.; Yang, Huanwang; Feng, Zukang; Baskaran, Kumaran; Berrisford, John M.; Hudson, Brian P.; Ikegawa, Yasuyo; Kobayashi, Naohiro; Lawson, Catherine L.; Mading, Steve; Mak, Lora; Mukhopadhyay, Abhik; Oldfield, Thomas J.; Patwardhan, Ardan; Peisach, Ezra; Sahni, Gaurav; Sekharan, Monica R.; Sen, Sanchayita; Shao, Chenghua; Smart, Oliver S.; Ulrich, Eldon L.; Yamashita, Reiko; Quesada, Martha; Young, Jasmine Y.; Nakamura, Haruki; Markley, John L.; Berman, Helen M.; Burley, Stephen K.; Velankar, Sameer; Kleywegt, Gerard J.Structure (Oxford, United Kingdom) (2017), 25 (12), 1916-1927CODEN: STRUE6; ISSN:0969-2126. (Elsevier Ltd.)The Worldwide PDB recently launched a deposition, biocuration, and validation tool: OneDep. At various stages of OneDep data processing, validation reports for three-dimensional structures of biol. macromols. are produced. These reports are based on recommendations of expert task forces representing crystallog., NMR, and cryoelectron microscopy communities. The reports provide useful metrics with which depositors can evaluate the quality of the exptl. data, the structural model, and the fit between them. The validation module is also available as a stand-alone web server and as a programmatically accessible web service. A growing no. of journals require the official wwPDB validation reports (produced at biocuration) to accompany manuscripts describing macromol. structures. Upon public release of the structure, the validation report becomes part of the public PDB archive. Geometric quality scores for proteins in the PDB archive have improved over the past decade.
- 75Schwarzinger, S.; Kroon, G. J.; Foss, T. R.; Wright, P. E.; Dyson, H. J. Random coil chemical shifts in acidic 8 M urea: implementation of random coil shift data in NMRView. J. Biomol. NMR 2000, 18 (1), 43– 48, DOI: 10.1023/A:1008386816521Google Scholar75https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXnvVGit7w%253D&md5=cf30cd19aa4420c807a0525789db9be1Random coil chemical shifts in acidic 8 M urea: implementation of random coil shift data in NMRViewSchwarzinger, Stephan; Kroon, Gerard J. A.; Foss, Ted R.; Wright, Peter E.; Dyson, H. JaneJournal of Biomolecular NMR (2000), 18 (1), 43-48CODEN: JBNME9; ISSN:0925-2738. (Kluwer Academic Publishers)Studies of proteins unfolded in acid or chem. denaturant can help in unraveling events during the earliest phases of protein folding. In order for meaningful comparisons to be made of residual structure in unfolded states, it is necessary to use random coil chem. shifts that are valid for the exptl. system under study. We present a set of random coil chem. shifts obtained for model peptides under exptl. conditions used in studies of denatured proteins. This new set, together with previously published data sets, has been incorporated into a software interface for NMRView, allowing selection of the random coil data set that fits the exptl. conditions best.
- 76Berendsen, H. J. C.; van der Spoel, D.; van Drunen, R. GROMACS: A message-passing parallel molecular dynamics implementation. Comput. Phys. Commun. 1995, 91 (1), 43– 56, DOI: 10.1016/0010-4655(95)00042-EGoogle Scholar76https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXps1Wrtr0%253D&md5=04d823aeab28ca374efb86839c705179GROMACS: A message-passing parallel molecular dynamics implementationBerendsen, H. J. C.; van der Spoel, D.; van Drunen, R.Computer Physics Communications (1995), 91 (1-3), 43-56CODEN: CPHCBZ; ISSN:0010-4655. (Elsevier)A parallel message-passing implementation of a mol. dynamics (MD) program that is useful for bio(macro)mols. in aq. environment is described. The software has been developed for a custom-designed 32-processor ring GROMACS (Groningen MAchine for Chem. Simulation) with communication to and from left and right neighbors, but can run on any parallel system onto which a a ring of processors can be mapped and which supports PVM-like block send and receive calls. The GROMACS software consists of a preprocessor, a parallel MD and energy minimization program that can use an arbitrary no. of processors (including one), an optional monitor, and several anal. tools. The programs are written in ANSI C and available by ftp (information: [email protected]). The functionality is based on the GROMOS (Groningen Mol. Simulation) package (van Gunsteren and Berendsen, 1987; BIOMOS B.V., Nijenborgh 4, 9747 AG Groningen). Conversion programs between GROMOS and GROMACS formats are included.The MD program can handle rectangular periodic boundary conditions with temp. and pressure scaling. The interactions that can be handled without modification are variable non-bonded pair interactions with Coulomb and Lennard-Jones or Buckingham potentials, using a twin-range cut-off based on charge groups, and fixed bonded interactions of either harmonic or constraint type for bonds and bond angles and either periodic or cosine power series interactions for dihedral angles. Special forces can be added to groups of particles (for non-equil. dynamics or for position restraining) or between particles (for distance restraints). The parallelism is based on particle decompn. Interprocessor communication is largely limited to position and force distribution over the ring once per time step.
- 77Jurrus, E.; Engel, D.; Star, K.; Monson, K.; Brandi, J.; Felberg, L. E.; Brookes, D. H.; Wilson, L.; Chen, J.; Liles, K. Improvements to the APBS biomolecular solvation software suite. Protein Sci. 2018, 27 (1), 112– 128, DOI: 10.1002/pro.3280Google Scholar77https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXhslSkt7vI&md5=99651d125e38f4a85d453fecf0f71652Improvements to the APBS biomolecular solvation software suiteJurrus, Elizabeth; Engel, Dave; Star, Keith; Monson, Kyle; Brandi, Juan; Felberg, Lisa E.; Brookes, David H.; Wilson, Leighton; Chen, Jiahui; Liles, Karina; Chun, Minju; Li, Peter; Gohara, David W.; Dolinsky, Todd; Konecny, Robert; Koes, David R.; Nielsen, Jens Erik; Head-Gordon, Teresa; Geng, Weihua; Krasny, Robert; Wei, Guo-Wei; Holst, Michael J.; McCammon, J. Andrew; Baker, Nathan A.Protein Science (2018), 27 (1), 112-128CODEN: PRCIEI; ISSN:1469-896X. (Wiley-Blackwell)The Adaptive Poisson-Boltzmann Solver (APBS) software was developed to solve the equations of continuum electrostatics for large biomol. assemblages that have provided impact in the study of a broad range of chem., biol., and biomedical applications. APBS addresses the three key technol. challenges for understanding solvation and electrostatics in biomedical applications: accurate and efficient models for biomol. solvation and electrostatics, robust and scalable software for applying those theories to biomol. systems, and mechanisms for sharing and analyzing biomol. electrostatics data in the scientific community. To address new research applications and advancing computational capabilities, we have continually updated APBS and its suite of accompanying software since its release in 2001. In this article, we discuss the models and capabilities that have recently been implemented within the APBS software package including a Poisson-Boltzmann anal. and a semi-anal. solver, an optimized boundary element solver, a geometry-based geometric flow solvation model, a graph theory-based algorithm for detg. pKa values, and an improved web-based visualization tool for viewing electrostatics.
- 78Martínez, L.; Andrade, R.; Birgin, E. G.; Martínez, J. M. PACKMOL: A package for building initial configurations for molecular dynamics simulations. J. Comput. Chem. 2009, 30 (13), 2157– 2164, DOI: 10.1002/jcc.21224Google Scholar78https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXptleqsb8%253D&md5=2a76255c873b866a26540f7e84496272PACKMOL: A package for building initial configurations for molecular dynamics simulationsMartinez, L.; Andrade, R.; Birgin, E. G.; Martinez, J. M.Journal of Computational Chemistry (2009), 30 (13), 2157-2164CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)Adequate initial configurations for mol. dynamics simulations consist of arrangements of mols. distributed in space in such a way to approx. represent the system's overall structure. In order that the simulations are not disrupted by large van der Waals repulsive interactions, atoms from different mols. must keep safe pairwise distances. Obtaining such a mol. arrangement can be considered a packing problem: Each type mol. must satisfy spatial constraints related to the geometry of the system, and the distance between atoms of different mols. must be greater than some specified tolerance. We have developed a code able to pack millions of atoms, grouped in arbitrarily complex mols., inside a variety of three-dimensional regions. The regions may be intersections of spheres, ellipses, cylinders, planes, or boxes. The user must provide only the structure of one mol. of each type and the geometrical constraints that each type of mol. must satisfy. Building complex mixts., interfaces, solvating biomols. in water, other solvents, or mixts. of solvents, is straightforward. In addn., different atoms belonging to the same mol. may also be restricted to different spatial regions, in such a way that more ordered mol. arrangements can be built, as micelles, lipid double-layers, etc. The packing time for state-of-the-art mol. dynamics systems varies from a few seconds to a few minutes in a personal computer. The input files are simple and currently compatible with PDB, Tinker, Molden, or Moldy coordinate files. The package is distributed as free software and can be downloaded from . © 2009 Wiley Periodicals, Inc. J Comput Chem, 2009.
- 79Rühle, V. Pressure coupling/barostats. 2008.Google ScholarThere is no corresponding record for this reference.
- 80Essmann, U.; Perera, L.; Berkowitz, M. L.; Darden, T.; Lee, H.; Pedersen, L. G. A smooth particle mesh Ewald method. J. Chem. Phys. 1995, 103 (19), 8577– 8593, DOI: 10.1063/1.470117Google Scholar80https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXptlehtrw%253D&md5=092a679dd3bee08da28df41e302383a7A smooth particle mesh Ewald methodEssmann, Ulrich; Perera, Lalith; Berkowitz, Max L.; Darden, Tom; Lee, Hsing; Pedersen, Lee G.Journal of Chemical Physics (1995), 103 (19), 8577-93CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)The previously developed particle mesh Ewald method is reformulated in terms of efficient B-spline interpolation of the structure factors. This reformulation allows a natural extension of the method to potentials of the form 1/rp with p ≥ 1. Furthermore, efficient calcn. of the virial tensor follows. Use of B-splines in the place of Lagrange interpolation leads to analytic gradients as well as a significant improvement in the accuracy. The authors demonstrate that arbitrary accuracy can be achieved, independent of system size N, at a cost that scales as N log(N). For biomol. systems with many thousands of atoms and this method permits the use of Ewald summation at a computational cost comparable to that of a simple truncation method of 10 Å or less.
- 81Lindorff-Larsen, K.; Piana, S.; Palmo, K.; Maragakis, P.; Klepeis, J. L.; Dror, R. O.; Shaw, D. E. Improved side-chain torsion potentials for the Amber ff99SB protein force field. Proteins 2010, 78, 1950– 1958, DOI: 10.1002/prot.22711Google Scholar81https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXkvFegtLo%253D&md5=447a9004026e2b93f0f7beff165daa09Improved side-chain torsion potentials for the Amber ff99SB protein force fieldLindorff-Larsen, Kresten; Piana, Stefano; Palmo, Kim; Maragakis, Paul; Klepeis, John L.; Dror, Ron O.; Shaw, David E.Proteins: Structure, Function, and Bioinformatics (2010), 78 (8), 1950-1958CODEN: PSFBAF ISSN:. (Wiley-Liss, Inc.)Recent advances in hardware and software have enabled increasingly long mol. dynamics (MD) simulations of biomols., exposing certain limitations in the accuracy of the force fields used for such simulations and spurring efforts to refine these force fields. Recent modifications to the Amber and CHARMM protein force fields, for example, have improved the backbone torsion potentials, remedying deficiencies in earlier versions. Here, the authors further advance simulation accuracy by improving the amino acid side-chain torsion potentials of the Amber ff99SB force field. First, the authors used simulations of model alpha-helical systems to identify the four residue types whose rotamer distribution differed the most from expectations based on Protein Data Bank statistics. Second, the authors optimized the side-chain torsion potentials of these residues to match new, high-level quantum-mech. calcns. Finally, the authors used microsecond-timescale MD simulations in explicit solvent to validate the resulting force field against a large set of exptl. NMR measurements that directly probe side-chain conformations. The new force field, which the authors have termed Amber ff99SB-ILDN, exhibits considerably better agreement with the NMR data. Proteins 2010. © 2010 Wiley-Liss, Inc.
- 82Mark, P.; Nilsson, L. Structure and Dynamics of the TIP3P, SPC, and SPC/E Water Models at 298 K. J. Phys. Chem. A 2001, 105 (43), 9954– 9960, DOI: 10.1021/jp003020wGoogle Scholar82https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXntlWrurs%253D&md5=fecd3db40210b04e8b2a933ea07b131eStructure and Dynamics of the TIP3P, SPC, and SPC/E Water Models at 298 KMark, Pekka; Nilsson, LennartJournal of Physical Chemistry A (2001), 105 (43), 9954-9960CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)Mol. dynamics simulations of five water models, the TIP3P (original and modified), SPC (original and refined), and SPC/E (original), were performed using the CHARMM mol. mechanics program. All simulations were carried out in the microcanonical NVE ensemble, using 901 water mols. in a cubic simulation cell furnished with periodic boundary conditions at 298 K. The SHAKE algorithm was used to keep water mols. rigid. Nanosecond trajectories were calcd. with all water models for high statistical accuracy. The characteristic self-diffusion coeffs. D and radial distribution functions, gOO, gOH, and gHH for all five water models were detd. and compared to exptl. data. The effects of velocity rescaling on the self-diffusion coeff. D were examd. All these empirical water models used in this study are similar by having three interaction sites, but the small differences in their pair potentials composed of Lennard-Jones (LJ) and Coulombic terms give significant differences in the calcd. self-diffusion coeffs., and in the height of the second peak of the radial distribution function gOO.
- 83de Vries, S. J.; van Dijk, M.; Bonvin, A. M. J. J. The HADDOCK web server for data-driven biomolecular docking. Nat. Protoc. 2010, 5 (5), 883– 897, DOI: 10.1038/nprot.2010.32Google Scholar83https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXltl2qsLw%253D&md5=dbf801d99065e8d2c3aedc016209838cThe HADDOCK web server for data-driven biomolecular dockingde Vries, Sjoerd J.; van Dijk, Marc; Bonvin, Alexandre M. J. J.Nature Protocols (2010), 5 (5), 883-897CODEN: NPARDW; ISSN:1750-2799. (Nature Publishing Group)Computational docking is the prediction or modeling of the three-dimensional structure of a biomol. complex, starting from the structures of the individual mols. in their free, unbound form. HADDOCK is a popular docking program that takes a data-driven approach to docking, with support for a wide range of exptl. data. Here the authors present the HADDOCK web server protocol, facilitating the modeling of biomol. complexes for a wide community. The main web interface is user-friendly, requiring only the structures of the individual components and a list of interacting residues as input. Addnl. web interfaces allow the more advanced user to exploit the full range of exptl. data supported by HADDOCK and to customize the docking process. The HADDOCK server has access to the resources of a dedicated cluster and of the e-NMR GRID infrastructure. Therefore, a typical docking run takes only a few minutes to prep. and a few hours to complete.
- 84Xue, L. C.; Rodrigues, J. P.; Kastritis, P. L.; Bonvin, A. M.; Vangone, A. PRODIGY: a web server for predicting the binding affinity of protein-protein complexes. Bioinformatics 2016, 32 (23), 3676– 3678, DOI: 10.1093/bioinformatics/btw514Google ScholarThere is no corresponding record for this reference.
- 85Schuck, P.; Zhao, H. Sedimentation Velocity Analytical Ultracentrifugation: Interacting Systems (1st ed.); CRC Press, 2017.Google ScholarThere is no corresponding record for this reference.
- 86Kyte, J.; Doolittle, R. F. A simple method for displaying the hydropathic character of a protein. J. Mol. Biol. 1982, 157 (1), 105– 132, DOI: 10.1016/0022-2836(82)90515-0Google Scholar86https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL38Xks1yjtro%253D&md5=ee67eb115939dfe56b2b2cae2c32dbd3A simple method for displaying the hydropathic character of a proteinKyte, Jack; Doolittle, Russell F.Journal of Molecular Biology (1982), 157 (1), 105-32CODEN: JMOBAK; ISSN:0022-2836.A computer program that progressively evaluates the hydrophilicity and hydrophobicity of a protein along its amino acid sequence was devised. A hydropathy scale takes into consideration the hydrophilic and hydrophobic properties of each of the 20 amino acid side chains. Correlation was demonstrated between the plotted values and known structures detd. by crystallog.
Cited By
This article has not yet been cited by other publications.
Article Views
Altmetric
Citations
Article Views are the COUNTER-compliant sum of full text article downloads since November 2008 (both PDF and HTML) across all institutions and individuals. These metrics are regularly updated to reflect usage leading up to the last few days.
Citations are the number of other articles citing this article, calculated by Crossref and updated daily. Find more information about Crossref citation counts.
The Altmetric Attention Score is a quantitative measure of the attention that a research article has received online. Clicking on the donut icon will load a page at altmetric.com with additional details about the score and the social media presence for the given article. Find more information on the Altmetric Attention Score and how the score is calculated.
Recommended Articles
Abstract
Figure 1
Figure 1. GNNV atomic model and cryo-EM structures in three different pH environments. (A) Atomic models of GNNV S-domain (PDB 4WIZ) and P-domain (PDB 4RFU). (B) Surface views of GNNV virus-like particles at pH 8.0 (left panel), 6.5 (middle panel), and 5.0 (right panel). The views are colored with a heatmap according to the local radius. (C) Conformational changes of the protrusion in response to the change in pH: 8.0 (left panel), 6.5 (middle panel), and 5.0 (right panel). The protrusions rotate clockwise and shift ∼5 Å toward the capsid shell as the pH decreases. (D) Enlarged views of pores (at 5- and 3-fold axis) in the capsid shell surface for different pH environments. The size of a pore at each pH condition is indicated. To highlight the pore at the 3-fold axis, the densities underneath are colored in light purple.
Figure 2
Figure 2. Effects of pH on GNNV-P in solution. (A) 1H–15N HSQC spectra of GNNV-P recorded at pH 7.0 (blue) and 5.0 (red). (B) Small box in panel (A) is enlarged. The six peaks, in green, at pH 7.0 are from Y279, L263, I222, I323, Q322, and R321, and the peaks, in red, at pH 5.0 are from E217 and I222. Y279, L263, I323, and Q322 disappear in the 1H–15N HSQC as the pH decreases. (C) Residue chemical shift perturbations (CSPs) by comparing 15N-labeled (pH 7.0) and 13C-, 15N-, and D-labeled (pH 5.0) HSQC spectra. The pH-sensitive Regions I (L263–Y279), II (W301–N303), and III (V312–V327) are highlighted in orange, yellow, and green, respectively. Residues with CSPs >2 standard deviations in pH-sensitive Regions I and III are highlighted in red and dark green, respectively. (D) pH-sensitive residues presented in panel (C) are mapped onto a surface representation of the GNNV-P crystal structure, (22) with the same color scheme as in panel (B).
Figure 3
Figure 3. Structure of GNNV-P determined in solution at neutral pH. (A) Sedimentation velocity analytical ultracentrifugation (SV AUC) data of GNNV-P obtained at pH 7.0 (black), 6.0 (blue), and 5.0 (red) fitted to a continuous sedimentation coefficient distribution c(s) model. Sedimentation coefficients and the molecular weight determined by SEDFIT are denoted above each peak. (B) Backbone (N, Cα, and C′) superimposition of the ensemble of 20 low-energy conformations. The N-terminal is colored blue, and the C-terminal is colored red. (C) Cartoon representation of the GNNV-P solution structure with β-strands and α-helices highlighted in red and blue, respectively. (D) Superimposition of the GNNV-P structure determined by NMR at neutral pH (pink) and by the X-ray crystal (PDB 4RFU, blue). (E) Schematic representation of GNNV-P secondary structures, as determined by NMR (pH 7.0) and X-ray crystallography (pH 6.5). β-strands are displayed as black arrows and α-helices as gray barrels, with bordering residues indicated.
Figure 4
Figure 4. MD-predicted structure of the GNNV-P oligomer at low pH. (A) Top and side view cartoon representations of the GNNV-P trimer formed under acidic pH conditions, as determined by molecular dynamics (MD) simulations. Chain A (blue), chain B (pink), and chain C (green). (B) Intermolecular interactions between neighboring GNNV-P units in the trimer. Residues at the trimeric interface are depicted as yellow sticks, and interactions are shown as dashed magenta lines. (C–E) Details of trimeric interface interactions between GNNV-P chains A and B, as boxed in panel (B).
Figure 5
Figure 5. GNNV-P undergoes a low-pH-induced conformational change. (A) Secondary structure propensities of GNNV-P at pH 7.0 and 5.0, calculated using secondary chemical ΔCα–ΔCβ shifts and plotted against the amino acid sequence. The Q320–P324 region showing an increased β-strand propensity at pH 5.0 is highlighted in gray. The secondary structures are shown above each chart, with arrows and cylinders representing β-strands and α-helices, respectively. (B) MD simulations revealed the formation of a short β-strand (comprising residues Q322–L324) within the F′–G′ loop at low pH. Hydrogen bonds between R321–L324 and D266–S264 were identified as stabilizing this region. (C) Superimposition of GNNV-P at pH 7.0 (NMR structure, blue) and 5.0 (MD model, red), showing the open and closed conformations of the F′–G′ loop, respectively. (D) 1H–15N HSQC spectra of uniform U–D-, 13C-, and 15N-labeled GNNV-P at pH 5.0. The regions with duplicate signals for residues 216–220 in the linker region are marked with black boxes. (E–G) Details of the uniform U–D-, 13C-, and 15N-labeled GNNV-P 1H–15N HSQC spectra highlighted by black boxes in panel (D). (H) Sequence of the N-terminal linker T214-P221 and a stick representation of the linker region with proline residues colored red and residues presenting duplicate NMR signals colored blue. (I) Schematic representation of cis–trans isomerization of an Xaa–Pro peptide bond. Proline residues can switch between the trans (blue) and cis (red) conformations.
Figure 6
Figure 6. Model of Neu5Ac-Lac binding to GNNV-P at neutral pH. (A) Surface representation of the GNNV-P monomer in complex with Neu5Ac-(α2,3)-Lac (green) and Neu5Ac-(α2,3)-Lac (yellow). (B) Details of Neu5Ac-(α2,3)-Lac binding to GNNV-P in the open pocket conformation. (C) Details of Neu5Ac-(α2,6)-Lac binding to GNNV-P in the open pocket conformation. In panels (B) and (C), residues interacting with Neu5Ac-Lac are shown as sticks, and selected contacts between GNNV-P and Neu5Ac-Lac are represented by pink dashed lines.
Figure 7
Figure 7. Model of GNNV interactions with host cell surface receptors. During infection, GNNV attaches to cells by interacting with the sialic acid moiety on the host cell surface and with cellular receptors (HSP70, HSP90ab1, or nectin-4). At weakly alkaline or neutral pH for host cell entry, protrusions on the GNNV surface are in extended and loose configuration, with the sialic acid binding pocket open, and linker region (red) and host-determining region (blue) accessible for interactions with cellular receptors. Acidification in late endosomes induces conformational change of the GNNV P-domain to result in resting/prone and compact protrusions. In this configuration, the sialic acid binding pocket is closed so that sialic acids are withdrawn to the tip of the protrusions. In addition, the accessibility of the linker region to a cellular receptor is also reprogrammed, leading to its potential GNNV detachment for endosomal escape. Note that the potential of the protrusion tip in directly interacting with the endosomal membrane may be altered by the pH-induced modification of the protrusion surface hydrophobicity distribution (Figure S20).
References
This article references 86 other publications.
- 1Baker, R. E.; Mahmud, A. S.; Miller, I. F.; Rajeev, M.; Rasambainarivo, F.; Rice, B. L.; Takahashi, S.; Tatem, A. J.; Wagner, C. E.; Wang, L.-F. Infectious disease in an era of global change. Nat. Rev. Microbiol. 2022, 20 (4), 193– 205, DOI: 10.1038/s41579-021-00639-z1https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB3MXit1Ciu77L&md5=e91900ca47ac5d7cf14a480ce6182ef9Infectious disease in an era of global changeBaker, Rachel E.; Mahmud, Ayesha S.; Miller, Ian F.; Rajeev, Malavika; Rasambainarivo, Fidisoa; Rice, Benjamin L.; Takahashi, Saki; Tatem, Andrew J.; Wagner, Caroline E.; Wang, Lin-Fa; Wesolowski, Amy; Metcalf, C. Jessica E.Nature Reviews Microbiology (2022), 20 (4), 193-205CODEN: NRMACK; ISSN:1740-1526. (Nature Portfolio)A review on. The twenty-first century has witnessed a wave of severe infectious disease outbreaks, not least the COVID-19 pandemic, which has had a devastating impact on lives and livelihoods around the globe. The 2003 severe acute respiratory syndrome coronavirus outbreak, the 2009 swine flu pandemic, the 2012 Middle East respiratory syndrome coronavirus outbreak, the 2013-2016 Ebola virus disease epidemic in West Africa and the 2015 Zika virus disease epidemic all resulted in substantial morbidity and mortality while spreading across borders to infect people in multiple countries. At the same time, the past few decades have ushered in an unprecedented era of technol., demog. and climatic change: airline flights have doubled since 2000, since 2007 more people live in urban areas than rural areas, population nos. continue to climb and climate change presents an escalating threat to society. In this Review, we consider the extent to which these recent global changes have increased the risk of infectious disease outbreaks, even as improved sanitation and access to health care have resulted in considerable progress worldwide.
- 2Doan, Q. K.; Vandeputte, M.; Chatain, B.; Morin, T.; Allal, F. Viral encephalopathy and retinopathy in aquaculture: a review. J. Fish Dis. 2017, 40 (5), 717– 742, DOI: 10.1111/jfd.12541There is no corresponding record for this reference.
- 3Cann, A. J. Replication of Viruses. In Encyclopedia of Virology (Third ed.); Mahy, B. W. J.; Van Regenmortel, M. H. V., Eds.; Academic Press, 2008; pp 406– 412.There is no corresponding record for this reference.
- 4Bruinsma, R. F.; Wuite, G. J. L.; Roos, W. H. Physics of viral dynamics. Nat. Rev. Phys. 2021, 3 (2), 76– 91, DOI: 10.1038/s42254-020-00267-1There is no corresponding record for this reference.
- 5Sherman, M. B.; Smith, H. Q.; Smith, T. J. The Dynamic Life of Virus Capsids. Viruses 2020, 12 (6), 618, DOI: 10.3390/v12060618There is no corresponding record for this reference.
- 6Zhang, Y.; Cui, Y.; Sun, J.; Zhou, Z. H. Multiple conformations of trimeric spikes visualized on a non-enveloped virus. Nat. Commun. 2022, 13 (1), 550 DOI: 10.1038/s41467-022-28114-0There is no corresponding record for this reference.
- 7Lynch, D. L.; Pavlova, A.; Fan, Z.; Gumbart, J. C. Understanding Virus Structure and Dynamics through Molecular Simulations. J. Chem. Theory Comput. 2023, 19 (11), 3025– 3036, DOI: 10.1021/acs.jctc.3c00116There is no corresponding record for this reference.
- 8Crane, M.; Hyatt, A. Viruses of fish: an overview of significant pathogens. Viruses 2011, 3 (11), 2025– 2046, DOI: 10.3390/v3112025There is no corresponding record for this reference.
- 9Mori, K.-I.; Nakai, T.; Muroga, K.; Arimoto, M.; Mushiake, K.; Furusawa, I. Properties of a new virus belonging to nodaviridae found in larval striped jack (Pseudocaranx dentex) with nervous necrosis. Virology 1992, 187 (1), 368– 371, DOI: 10.1016/0042-6822(92)90329-N9https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADyaK387kt1Ghug%253D%253D&md5=2e07e4427b2baf88353ad92b9de7012dProperties of a new virus belonging to nodaviridae found in larval striped jack (Pseudocaranx dentex) with nervous necrosisMori K; Nakai T; Muroga K; Arimoto M; Mushiake K; Furusawa IVirology (1992), 187 (1), 368-71 ISSN:0042-6822.Spherical virus particles were purified from larval striped jack (Pseudocaranx dentex) with nervous necrosis. The virus consists of nonenveloped particles, about 25 nm in diameter, and contains two single-stranded, positive-sense RNA molecules with molecular weights of 1.01 x 10(6)Da (RNA 1) and 0.49 x 10(6)Da (RNA 2), respectively. The RNAs do not have poly(A) sequences at the 3' terminus. Virus structural proteins consist of two proteins with molecular weights of 42 and 40 kDa. When translated into cell-free extracts of rabbit reticulocytes, RNA 1 directed the synthesis of the 1a protein (100 kDa), whereas RNA 2 synthesized the 2a protein (42 kDa), which is probably the coat protein of the virus, and a polypeptide of 40 kDa which appears to be the processed form of the 42-kDa protein. Under electron microscopic observation, the virus particles were found in the tissues of the central nervous system of the affected larval striped jack. From morphological and biochemical properties of the virus, we identified this virus as a new member of the family of Nodaviridae and designated it striped jack nervous necrosis virus.
- 10Bandín, I.; Souto, S. Betanodavirus and VER Disease: A 30-year Research Review. Pathogens 2020, 9 (2), 106, DOI: 10.3390/pathogens9020106There is no corresponding record for this reference.
- 11Venter, P. A.; Schneemann, A. Nodaviruses. In Encyclopedia of Virology (Third ed.); Mahy, B. W. J.; Van Regenmortel, M. H. V., Eds.; Academic Press, 2008; pp 430– 438.There is no corresponding record for this reference.
- 12Tang, L.; Lin, C.-S.; Krishna, N. K.; Yeager, M.; Schneemann, A.; Johnson, J. E. Virus-like particles of a fish nodavirus display a capsid subunit domain organization different from that of insect nodaviruses. J. Virol. 2002, 76 (12), 6370– 6375, DOI: 10.1128/JVI.76.12.6370-6375.2002There is no corresponding record for this reference.
- 13Iwamoto, T.; Mise, K.; Takeda, A.; Okinaka, Y.; Mori, K.-I.; Arimoto, M.; Okuno, T.; Nakai, T. Characterization of Striped jack nervous necrosis virus subgenomic RNA3 and biological activities of its encoded protein B2. J. Gen. Virol. 2005, 86, 2807– 2816, DOI: 10.1099/vir.0.80902-0There is no corresponding record for this reference.
- 14Liu, W.; Hsu, C.-H.; Hong, Y.-R.; Wu, S.-C.; Wang, C.-H.; Wu, Y.-M.; Chao, C.-B.; Lin, C.-S. Early endocytosis pathways in SSN-1 cells infected by dragon grouper nervous necrosis virus. J. Gen. Virol. 2005, 86 (9), 2553– 2561, DOI: 10.1099/vir.0.81021-0There is no corresponding record for this reference.
- 15Adachi, K.; Ichinose, T.; Takizawa, N.; Watanabe, K.; Kitazato, K.; Kobayashi, N. Inhibition of betanodavirus infection by inhibitors of endosomal acidification. Arch. Virol. 2007, 152 (12), 2217– 2224, DOI: 10.1007/s00705-007-1061-7There is no corresponding record for this reference.
- 16Huang, R.; Zhu, G.; Zhang, J.; Lai, Y.; Xu, Y.; He, J.; Xie, J. Betanodavirus-like particles enter host cells via clathrin-mediated endocytosis in a cholesterol-, pH- and cytoskeleton-dependent manner. Vet. Res. 2017, 48 (1), 8 DOI: 10.1186/s13567-017-0412-yThere is no corresponding record for this reference.
- 17Kumar, C. S.; Dey, D.; Ghosh, S.; Banerjee, M. Breach: Host Membrane Penetration and Entry by Nonenveloped Viruses. Trends Microbiol. 2018, 26 (6), 525– 537, DOI: 10.1016/j.tim.2017.09.010There is no corresponding record for this reference.
- 18Mercer, J.; Schelhaas, M.; Helenius, A. Virus Entry by Endocytosis. Annu. Rev. Biochem. 2010, 79 (1), 803– 833, DOI: 10.1146/annurev-biochem-060208-10462618https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXpslSht7c%253D&md5=e083b9661ae42c15824b304f13c8e6fcVirus entry by endocytosisMercer, Jason; Schelhaas, Mario; Helenius, AriAnnual Review of Biochemistry (2010), 79 (), 803-833CODEN: ARBOAW; ISSN:0066-4154. (Annual Reviews Inc.)A review. Although viruses are simple in structure and compn., their interactions with host cells are complex. Merely to gain entry, animal viruses make use of a repertoire of cellular processes that involve hundreds of cellular proteins. Although some viruses have the capacity to penetrate into the cytosol directly through the plasma membrane, most depend on endocytic uptake, vesicular transport through the cytoplasm, and delivery to endosomes and other intracellular organelles. The internalization may involve clathrin-mediated endocytosis (CME), macropinocytosis, caveolar/lipid raft-mediated endocytosis, or a variety of other still poorly characterized mechanisms. This review focuses on the cell biol. of virus entry and the different strategies and endocytic mechanisms used by animal viruses.
- 19Pletan, M. L.; Tsai, B. Non-enveloped virus membrane penetration: New advances leading to new insights. PLoS Pathog. 2022, 18 (12), e1010948 DOI: 10.1371/journal.ppat.1010948There is no corresponding record for this reference.
- 20Odegard, A.; Banerjee, M.; Johnson, J. E. Flock House Virus: A Model System for Understanding Non-Enveloped Virus Entry and Membrane Penetration. In Cell Entry by Non-Enveloped Viruses; Johnson, J. E., Ed.; Springer Berlin Heidelberg, 2010; pp 1– 22.There is no corresponding record for this reference.
- 21Odegard, A. L.; Kwan, M. H.; Walukiewicz, H. E.; Banerjee, M.; Schneemann, A.; Johnson, J. E. Low endocytic pH and capsid protein autocleavage are critical components of Flock House virus cell entry. J. Virol. 2009, 83 (17), 8628– 8637, DOI: 10.1128/JVI.00873-0921https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXhtFahurrJ&md5=ca9a32cff587b467ac1302d995eff723Low endocytic pH and capsid protein autocleavage are critical components of flock house virus cell entryOdegard, Amy L.; Kwan, Maggie H.; Walukiewicz, Hanna E.; Banerjee, Manidipa; Schneemann, Anette; Johnson, John E.Journal of Virology (2009), 83 (17), 8628-8637CODEN: JOVIAM; ISSN:0022-538X. (American Society for Microbiology)The process by which nonenveloped viruses cross cell membranes during host cell entry remains poorly defined; however, common themes are emerging. Here, we use correlated in vivo and in vitro studies to understand the mechanism of Flock House virus (FHV) entry and membrane penetration. We demonstrate that low endocytic pH is required for FHV infection, that exposure to acidic pH promotes FHV-mediated disruption of model membranes (liposomes), and particles exposed to low pH in vitro exhibit increased hydrophobicity. In addn., FHV particles perturbed by heating displayed a marked increase in liposome disruption, indicating that membrane-active regions of the capsid are exposed or released under these conditions. We also provide evidence that autoproteolytic cleavage, to generate the lipophilic γ peptide (4.4 kDa), is required for membrane penetration. Mutant, cleavage-defective particles failed to mediate liposome lysis, regardless of pH or heat treatment, suggesting that these particles are not able to expose or release the requisite membrane-active regions of the capsid, namely, the γ peptides. Based on these results, we propose an updated model for FHV entry in which (i) the virus enters the host cell by endocytosis, (ii) low pH within the endocytic pathway triggers the irreversible exposure or release of γ peptides from the virus particle, and (iii) the exposed/released γ peptides disrupt the endosomal membrane, facilitating translocation of viral RNA into the cytoplasm.
- 22Chen, N.-C.; Yoshimura, M.; Guan, H.-H.; Wang, T.-Y.; Misumi, Y.; Lin, C.-C.; Chuankhayan, P.; Nakagawa, A.; Chan, S. I.; Tsukihara, T. Crystal Structures of a Piscine Betanodavirus: Mechanisms of Capsid Assembly and Viral Infection. PLoS Pathog. 2015, 11 (10), e1005203, DOI: 10.1371/journal.ppat.1005203There is no corresponding record for this reference.
- 23Gye, H. J.; Nishizawa, T. Sites responsible for infectivity and antigenicity on nervous necrosis virus (NNV) appear to be distinct. Sci. Rep. 2021, 11 (1), 3608 DOI: 10.1038/s41598-021-83078-3There is no corresponding record for this reference.
- 24Gye, H. J.; Park, M.-J.; Kim, W.-S.; Oh, M.-J.; Nishizawa, T. Heat-denaturation of conformational structures on nervous necrosis virus for generating neutralization antibodies. Aquaculture 2018, 484, 65– 70, DOI: 10.1016/j.aquaculture.2017.10.034There is no corresponding record for this reference.
- 25Nishizawa, T.; Mori, K.-i.; Furuhashi, M.; Nakai, T.; Furusawa, I.; Muroga, K. Comparison of the coat protein genes of five fish nodaviruses, the causative agents of viral nervous necrosis in marine fish. J. Gen. Virol. 1995, 76 (7), 1563– 1569, DOI: 10.1099/0022-1317-76-7-1563There is no corresponding record for this reference.
- 26Iwamoto, T.; Okinaka, Y.; Mise, K.; Mori, K.-I.; Arimoto, M.; Okuno, T.; Nakai, T. Identification of Host-Specificity Determinants in Betanodaviruses by Using Reassortants between Striped Jack Nervous Necrosis Virus and Sevenband Grouper Nervous Necrosis Virus. J. Virol. 2004, 78 (3), 1256– 1262, DOI: 10.1128/JVI.78.3.1256-1262.2004There is no corresponding record for this reference.
- 27Moreno, P.; Souto, S.; Leiva-Rebollo, R.; Borrego, J. J.; Bandín, I.; Alonso, M. C. Capsid amino acids at positions 247 and 270 are involved in the virulence of betanodaviruses to European sea bass. Sci. Rep. 2019, 9 (1), 14068 DOI: 10.1038/s41598-019-50622-1There is no corresponding record for this reference.
- 28Nishizawa, T.; Lee, H. S.; Gye, H. J. Pocket structures of surface protrusions shared among serologically distinct nervous necrosis viruses (NNVs) were predicted in silico to bind to sialylated N-glycans, a host cellular receptor. Aquaculture 2023, 565, 739157, DOI: 10.1016/j.aquaculture.2022.739157There is no corresponding record for this reference.
- 29Wang, C. H.; Chen, D. H.; Huang, S. H.; Wu, Y. M.; Chen, Y. Y.; Hwu, Y.; Bushnell, D.; Kornberg, R.; Chang, W. H. Sub-3 Å Cryo-EM Structures of Necrosis Virus Particles via the Use of Multipurpose TEM with Electron Counting Camera. Int. J. Mol. Sci. 2021, 22 (13), 6859, DOI: 10.3390/ijms22136859There is no corresponding record for this reference.
- 30Chang, W. H.; Huang, S. H.; Lin, H. H.; Chung, S. C.; Tu, I. P. Cryo-EM Analyses Permit Visualization of Structural Polymorphism of Biological Macromolecules. Front. Bioinform. 2021, 1, 788308, DOI: 10.3389/fbinf.2021.788308There is no corresponding record for this reference.
- 31Zhang, K.; Julius, D.; Cheng, Y. Structural snapshots of TRPV1 reveal mechanism of polymodal functionality. Cell 2021, 184 (20), 5138– 5150, DOI: 10.1016/j.cell.2021.08.012There is no corresponding record for this reference.
- 32Chen, C.-Y.; Chang, Y.-C.; Lin, B.-L.; Huang, C.-H.; Tsai, M.-D. Temperature-Resolved Cryo-EM Uncovers Structural Bases of Temperature-Dependent Enzyme Functions. J. Am. Chem. Soc. 2019, 141 (51), 19983– 19987, DOI: 10.1021/jacs.9b1068732https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXitlygt7zO&md5=3866b861e15dbebe8e5657828643cbdaTemperature-Resolved Cryo-EM Uncovers Structural Bases of Temperature-Dependent Enzyme FunctionsChen, Chin-Yu; Chang, Yuan-Chih; Lin, Bo-Lin; Huang, Chun-Hsiang; Tsai, Ming-DawJournal of the American Chemical Society (2019), 141 (51), 19983-19987CODEN: JACSAT; ISSN:0002-7863. (American Chemical Society)Protein functions are temp.-dependent, but protein structures are usually solved at a single (often low) temp. because of limitations on the conditions of crystal growth or protein vitrification. Here we demonstrate the feasibility of solving cryo-EM structures of proteins vitrified at high temps., solve 12 structures of an archaeal ketol-acid reductoisomerase (KARI) vitrified at 4-70 °C, and show that structures of both the Mg2+ form (KARI:2Mg2+) and its ternary complex (KARI:2Mg2+:NADH:inhibitor) are temp.-dependent in correlation with the temp. dependence of enzyme activity. Furthermore, structural analyses led to dissection of the induced-fit mechanism into ligand-induced and temp.-induced effects and to capture of temp.-resolved intermediates of the temp.-induced conformational change. The results also suggest that it is preferable to solve cryo-EM structures of protein complexes at functional temps. These studies should greatly expand the landscapes of protein structure-function relationships and enhance the mechanistic anal. of enzymic functions.
- 33Xie, J.; Li, K.; Gao, Y.; Huang, R.; Lai, Y.; Shi, Y.; Yang, S.; Zhu, G.; Zhang, Q.; He, J. Structural analysis and insertion study reveal the ideal sites for surface displaying foreign peptides on a betanodavirus-like particle. Vet. Res. 2016, 47 (1), 16, DOI: 10.1186/s13567-015-0294-9There is no corresponding record for this reference.
- 34Štěrbová, P.; Wu, D.; Lou, Y.-C.; Wang, C.-H.; Chang, W.-H.; Tzou, D.-L. M. NMR assignments of protrusion domain of capsid protein from dragon grouper nervous necrosis virus. Biomol. NMR Assignments 2020, 14 (1), 63– 66, DOI: 10.1007/s12104-019-09921-xThere is no corresponding record for this reference.
- 35Walters, K. J.; Ferentz, A. E.; Hare, B. J.; Hidalgo, P.; Jasanoff, A.; Matsuo, H.; Wagner, G. Characterizing protein-protein complexes and oligomers by nuclear magnetic resonance spectroscopy. In Methods in Enzymology; Elsevier, 2001; Vol. 339, pp 238– 258. DOI: 10.1016/S0076-6879(01)39316-3 .There is no corresponding record for this reference.
- 36Williamson, M. P. Using chemical shift perturbation to characterise ligand binding. Prog. Nucl. Magn. Reson. Spectrosc. 2013, 73, 1– 16, DOI: 10.1016/j.pnmrs.2013.02.00136https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3sXht1yrsbzE&md5=74fc06bb4f4b3df369619bbf730016ceUsing chemical shift perturbation to characterise ligand bindingWilliamson, Mike P.Progress in Nuclear Magnetic Resonance Spectroscopy (2013), 73 (), 1-16CODEN: PNMRAT; ISSN:0079-6565. (Elsevier B.V.)A review. Chem. shift perturbation (CSP, chem. shift mapping or complexation-induced changes in chem. shift, CIS) follows changes in the chem. shifts of a protein when a ligand is added, and uses these to det. the location of the binding site, the affinity of the ligand, and/or possibly the structure of the complex. A key factor in detg. the appearance of spectra during a titrn. is the exchange rate between free and bound, or more specifically the off-rate koff. When koff is greater than the chem. shift difference between free and bound, which typically equates to an affinity Kd weaker than about 3 μM, then exchange is fast on the chem. shift timescale. Under these circumstances, the obsd. shift is the population-weighted av. of free and bound, which allows Kd to be detd. from measurement of peak positions, provided the measurements are made appropriately. 1H shifts are influenced to a large extent by through-space interactions, whereas 13Cα and 13Cβ shifts are influenced more by through-bond effects. 15N and 13C' shifts are influenced both by through-bond and by through-space (hydrogen bonding) interactions. For detg. the location of a bound ligand on the basis of shift change, the most appropriate method is therefore usually to measure 15N HSQC spectra, calc. the geometrical distance moved by the peak, weighting 15N shifts by a factor of about 0.14 compared to 1H shifts, and select those residues for which the weighted shift change is larger than the std. deviation of the shift for all residues. Other methods are discussed, in particular the measurement of 13CH3 signals. Slow to intermediate exchange rates lead to line broadening, and make Kd values very difficult to obtain. There is no good way to distinguish changes in chem. shift due to direct binding of the ligand from changes in chem. shift due to allosteric change. Ligand binding at multiple sites can often be characterized, by simultaneous fitting of many measured shift changes, or more simply by adding substoichiometric amts. of ligand. The chem. shift changes can be used as restraints for docking ligand onto protein. By use of quant. calcns. of ligand-induced chem. shift changes, it is becoming possible to det. not just the position but also the orientation of ligands.
- 37Krissinel, E.; Henrick, K. Inference of Macromolecular Assemblies from Crystalline State. J. Mol. Biol. 2007, 372 (3), 774– 797, DOI: 10.1016/j.jmb.2007.05.02237https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD2sXpvFGktb8%253D&md5=a5c764cfc7dc129f53ddc31ef9d475faInference of Macromolecular Assemblies from Crystalline StateKrissinel, Evgeny; Henrick, KimJournal of Molecular Biology (2007), 372 (3), 774-797CODEN: JMOBAK; ISSN:0022-2836. (Elsevier Ltd.)The authors discuss basic phys.-chem. principles underlying the formation of stable macromol. complexes, which in many cases are likely to be the biol. units performing a certain physiol. function. The authors also consider available theor. approaches to the calcn. of macromol. affinity and entropy of complexation. The latter is shown to play an important role and make a major effect on complex size and symmetry. The authors develop a new method, based on chem. thermodn., for automatic detection of macromol. assemblies in the Protein Data Bank (PDB) entries that are the results of x-ray diffraction expts. As found, biol. units may be recovered at 80-90% success rate, which makes x-ray crystallog. an important source of exptl. data on macromol. complexes and protein-protein interactions. The method is implemented as a public WWW service (http://www.ebi.ac.uk/msd-srv/prot_int/pistart.html).
- 38Kampmann, T.; Mueller, D. S.; Mark, A. E.; Young, P. R.; Kobe, B. The Role of histidine residues in low-pH-mediated viral membrane fusion. Structure 2006, 14 (10), 1481– 1487, DOI: 10.1016/j.str.2006.07.01138https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28XhtVCrtrnN&md5=ffb8d25828217e82747ef2838f53f028The Role of Histidine Residues in Low-pH-Mediated Viral Membrane FusionKampmann, Thorsten; Mueller, Daniela S.; Mark, Alan E.; Young, Paul R.; Kobe, BostjanStructure (Cambridge, MA, United States) (2006), 14 (10), 1481-1487CODEN: STRUE6; ISSN:0969-2126. (Cell Press)A central event in the invasion of a host cell by an enveloped virus is the fusion of viral and cell membranes. For many viruses, membrane fusion is driven by specific viral surface proteins that undergo large-scale conformational rearrangements, triggered by exposure to low pH in the endosome upon internalization. Here, we present evidence suggesting that in both class I (helical hairpin proteins) and class II (β-structure-rich proteins) pH-dependent fusion proteins the protonation of specific histidine residues triggers fusion via an analogous mol. mechanism. These histidines are located in the vicinity of pos. charged residues in the prefusion conformation, and they subsequently form salt bridges with neg. charged residues in the postfusion conformation. The mol. surfaces involved in the corresponding structural rearrangements leading to fusion are highly conserved and thus might provide a suitable common target for the design of antivirals, which could be active against a diverse range of pathogenic viruses.
- 39The PyMOL Molecular Graphics System, Version 1.8, Schrodinger, LLC, 2015.There is no corresponding record for this reference.
- 40Wedemeyer, W. J.; Welker, E.; Scheraga, H. A. Proline cis-trans isomerization and protein folding. Biochemistry 2002, 41 (50), 14637– 14644, DOI: 10.1021/bi020574b40https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD38XotlOmsLw%253D&md5=f05d94502166b940f92d215d49352b6fProline Cis-Trans Isomerization and Protein FoldingWedemeyer, William J.; Welker, Ervin; Scheraga, Harold A.Biochemistry (2002), 41 (50), 14637-14644CODEN: BICHAW; ISSN:0006-2960. (American Chemical Society)A review. Proline cis-trans isomerization plays a key role in the rate-detg. steps of protein folding. The energetic origin of this isomerization process is summarized, and the folding and unfolding of disulfide-intact bovine pancreatic RNase A is used as an example to illustrate the kinetics and structural features of conformational changes from the heterogeneous unfolded state (consisting of cis and trans isomers of X-Pro peptide groups) to the native structure in which only one set of proline isomers is present.
- 41Shen, Y.; Bax, A. Prediction of Xaa-Pro peptide bond conformation from sequence and chemical shifts. J. Biomol. NMR 2010, 46 (3), 199– 204, DOI: 10.1007/s10858-009-9395-y41https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXivV2ku70%253D&md5=30d471955095ff1da608e18b1d3321d4Prediction of Xaa-Pro peptide bond conformation from sequence and chemical shiftsShen, Yang; Bax, AdJournal of Biomolecular NMR (2010), 46 (3), 199-204CODEN: JBNME9; ISSN:0925-2738. (Springer)The authors present a program, named Promega, to predict the Xaa-Pro peptide bond conformation on the basis of backbone chem. shifts and the amino acid sequence. Using a chem. shift database of proteins of known structure together with the PDB-extd. amino acid preference of cis Xaa-Pro peptide bonds, a cis/trans probability score is calcd. from the backbone and 13Cβ chem. shifts of the proline and its neighboring residues. For an arbitrary no. of input chem. shifts, which may include Pro-13Cγ, Promega calcs. the statistical probability that a Xaa-Pro peptide bond is cis. Besides its potential as a validation tool, Promega is particularly useful for studies of larger proteins where Pro-13Cγ assignments can be challenging, and for on-going efforts to det. protein structures exclusively on the basis of backbone and 13Cβ chem. shifts.
- 42Gye, H. J.; Nishizawa, T. Analysis of sialylated N-linked glycans on fish cell lines permissive to nervous necrosis virus for predicting cellular receptors of the virus. Aquaculture 2022, 555, 738198 DOI: 10.1016/j.aquaculture.2022.738198There is no corresponding record for this reference.
- 43Rodríguez, Y.; Cardoze, S. M.; Obineche, O. W.; Melo, C.; Persaud, A.; Fernández Romero, J. A. Small Molecules Targeting SARS-CoV-2 Spike Glycoprotein Receptor-Binding Domain. ACS Omega 2022, 7 (33), 28779– 28789, DOI: 10.1021/acsomega.2c0084443https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BB38XitVKktrbO&md5=4c573b63f6c38170fbe0bff8533dd57bSmall Molecules Targeting SARS-CoV-2 Spike Glycoprotein Receptor-Binding DomainRodriguez, Yoel; Cardoze, Scarlet Martinez; Obineche, Onyinyechi W.; Melo, Claudia; Persaud, Ashanna; Fernandez Romero, Jose A.ACS Omega (2022), 7 (33), 28779-28789CODEN: ACSODF; ISSN:2470-1343. (American Chemical Society)The severe acute respiratory syndrome coronavirus 2 (SARS-CoV-2) has caused the coronavirus disease 2019 (COVID-19) pandemic. Several variants of SARS-CoV-2 have emerged worldwide. These variants show different transmissibility infectivity due to mutations in the viral spike (S) glycoprotein that interacts with the human angiotensin-converting enzyme 2 (hACE2) receptor and facilitates viral entry into target cells. Despite the effective SARS-CoV-2 vaccines, we still need to identify selective antivirals, and the S glycoprotein is a key target to neutralize the virus. We hypothesize that small mols. could disrupt the interaction of S glycoprotein with hACE2 and inhibit viral entry. We analyzed the S glycoprotein-hACE2 complex structure (PDB: 7DF4) and created models for different viral variants using visual mol. dynamics (VMD) and mol. operating environment (MOE) programs. Moreover, we started the hits search by performing structure-based mol. docking virtual screening of com. available small mols. against S glycoprotein models using OEDocking FRED-4.0.0.0 software. The FRED-4.0.0.0 Chemguass4 scoring function was used to rank the small mols. based on their affinities. The best candidate compds. were purchased and tested using a std. SARS-CoV-2 pseudotyped cell-based bioassay to investigate their antiviral activity. Three of these compds., alone or in combination, showed antiviral selectivity. These small mols. may lead to an effective antiviral treatment or serve as probes to better understand the biol. of SARS-CoV-2.
- 44Chang, J.-S.; Chi, S.-C. GHSC70 is involved in the cellular entry of nervous necrosis virus. J. Virol. 2015, 89 (1), 61– 70, DOI: 10.1128/JVI.02523-14There is no corresponding record for this reference.
- 45Zhang, W.; Jia, K.; Jia, P.; Xiang, Y.; Lu, X.; Liu, W.; Yi, M. Marine medaka heat shock protein 90ab1 is a receptor for red-spotted grouper nervous necrosis virus and promotes virus internalization through clathrin-mediated endocytosis. PLoS Pathog. 2020, 16 (7), e1008668 DOI: 10.1371/journal.ppat.1008668There is no corresponding record for this reference.
- 46Krishnan, R.; Qadiri, S. S. N.; Oh, M.-J. Functional characterization of seven-band grouper immunoglobulin like cell adhesion molecule, Nectin4 as a cellular receptor for nervous necrosis virus. Fish Shellfish Immunol. 2019, 93, 720– 725, DOI: 10.1016/j.fsi.2019.08.019There is no corresponding record for this reference.
- 47Ito, Y.; Okinaka, Y.; Mori, K. I.; Sugaya, T.; Nishioka, T.; Oka, M.; Nakai, T. Variable region of betanodavirus RNA2 is sufficient to determine host specificity. Dis. Aquat. Org. 2008, 79 (3), 199– 205, DOI: 10.3354/dao01906There is no corresponding record for this reference.
- 48Staring, J.; Raaben, M.; Brummelkamp, T. R. Viral escape from endosomes and host detection at a glance. J. Cell Sci. 2018, 131 (15), jcs216259 DOI: 10.1242/jcs.216259There is no corresponding record for this reference.
- 49Song, C.; Takai-Todaka, R.; Miki, M.; Haga, K.; Fujimoto, A.; Ishiyama, R.; Oikawa, K.; Yokoyama, M.; Miyazaki, N.; Iwasaki, K. Dynamic rotation of the protruding domain enhances the infectivity of norovirus. PLoS Pathog. 2020, 16 (7), e1008619 DOI: 10.1371/journal.ppat.1008619There is no corresponding record for this reference.
- 50Hu, L.; Salmen, W.; Chen, R.; Zhou, Y.; Neill, F.; Crowe, J. E.; Atmar, R. L.; Estes, M. K.; Prasad, B. V. V. Atomic structure of the predominant GII.4 human norovirus capsid reveals novel stability and plasticity. Nat. Commun. 2022, 13 (1), 1241 DOI: 10.1038/s41467-022-28757-zThere is no corresponding record for this reference.
- 51Goto, Y.; Takahashi, N.; Fink, A. L. (1990). Mechanism of acid-induced folding of proteins. Biochemistry 1990, 29 (14), 3480– 3488, DOI: 10.1021/bi00466a00951https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK3cXhsFeqs7w%253D&md5=48e21a0012913f7c73f4748b2c62a142Mechanism of acid-induced folding of proteinsGoto, Yuji; Takahashi, Nobuaki; Fink, Anthony L.Biochemistry (1990), 29 (14), 3480-8CODEN: BICHAW; ISSN:0006-2960.It was previously shown that β-lactamase, cytochrome c (I), and apomyoglobin (II) are maximally unfolded at pH 2 under conditions of low ionic strength, but that a further decrease in pH, by increasing the concn. of HCl, refolds the proteins to the A state with properties similar to those of a molten globule state. To understand the mechanism of acid-induced refolding of protein structure, the effects of various strong acids and their neutral salts on the acid-unfolded states of ferri-I and II were studied. The conformational transition of I was monitored at 20° by using changes in the far-UV CD and in the Soret absorption at 394 nm, and that of II was monitored by changes in the far-UV CD. Various strong acids (i.e., H2SO4, HClO4, HNO3, TCA, and trifluoroacetic acid) refolded acid-unfolded I and II to the A states as was the case with HCl. For both proteins, neutral salts of these acids caused similar conformational transitions, confirming that the anions are responsible for bringing about the transition. The order of effectiveness of anions was ferricyanide > ferrocyanide > SO42- > TCA- > SCN- > ClO4- > I- > NO3- > trifluoroacetate > Br- > Cl-. This series was similar to the electroselectivity series of anions toward the anion-exchange resins, showing that preferential binding of anions to the A states causes the conformational transitions.
- 52Güthe, S.; Kapinos, L.; Möglich, A.; Meier, S.; Grzesiek, S.; Kiefhaber, T. Very fast folding and association of a trimerization domain from bacteriophage T4 fibritin. J. Mol. Biol. 2004, 337 (4), 905– 915, DOI: 10.1016/j.jmb.2004.02.020There is no corresponding record for this reference.
- 53So, M.; Hall, D.; Goto, Y. Revisiting supersaturation as a factor determining amyloid fibrillation. Curr. Opin. Struct. Biol. 2016, 36, 32– 39, DOI: 10.1016/j.sbi.2015.11.00953https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2MXitVSkurjP&md5=01c17457e922fbefc12715f0bf000e1dRevisiting supersaturation as a factor determining amyloid fibrillationSo, Masatomo; Hall, Damien; Goto, YujiCurrent Opinion in Structural Biology (2016), 36 (), 32-39CODEN: COSBEF; ISSN:0959-440X. (Elsevier Ltd.)Amyloid fibrils involved in various diseases are formed by a nucleation-growth mechanism, similar to the crystn. of solutes from soln. Soly. and supersatn. are two of the most important factors detg. crystn. of solutes. Moreover, crystn. competes with glass formation in which solutes collapse into amorphous aggregates. Recent studies on the formation of amyloid fibrils and amorphous aggregates indicate that the partition between distinct types of aggregates can be rationally explained by a kinetic and thermodn. competition between them. Understanding the role of supersatn. in detg. aggregation-based phase transitions of denatured proteins provides an important complementary point of view to structural studies of protein aggregates.
- 54Shih, T. C.; Ho, L. P.; Chou, H. Y.; Wu, J. L.; Pai, T. W. Comprehensive Linear Epitope Prediction System for Host Specificity in Nodaviridae. Viruses 2022, 14 (7), 1357, DOI: 10.3390/v14071357There is no corresponding record for this reference.
- 55Zhang, Z.; Xing, J.; Tang, X.; Sheng, X.; Chi, H.; Zhan, W. Development and characterization of monoclonal antibodies against red-spotted grouper nervous necrosis virus and their neutralizing potency in vitro. Aquaculture 2022, 560, 738562 DOI: 10.1016/j.aquaculture.2022.738562There is no corresponding record for this reference.
- 56Huang, S.; Wu, Y.; Su, L.; Su, T.; Zhou, Q.; Zhang, J.; Zhao, Z.; Weng, S.; He, J.; Xie, J. A single-chain variable fragment antibody exerts anti-nervous necrosis virus activity by irreversible binding. Aquaculture 2022, 552, 738001 DOI: 10.1016/j.aquaculture.2022.738001There is no corresponding record for this reference.
- 57Graham, B. S.; Gilman, M. S. A.; McLellan, J. S. Structure-Based Vaccine Antigen Design. Annu. Rev. Med. 2019, 70, 91– 104, DOI: 10.1146/annurev-med-121217-09423457https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC1MXhvF2ltrw%253D&md5=ccbef186198ab5f51fe6aa228b440734Structure-Based Vaccine Antigen DesignGraham, Barney S.; Gilman, Morgan S. A.; McLellan, Jason S.Annual Review of Medicine (2019), 70 (), 91-104CODEN: ARMCAH; ISSN:0066-4219. (Annual Reviews)A review. Enabled by new approaches for rapid identification and selection of human monoclonal antibodies, at.-level structural information for viral surface proteins, and capacity for precision engineering of protein immunogens and self-assembling nanoparticles, a new era of antigen design and display options has evolved. While HIV-1 vaccine development has been a driving force behind these technologies and concepts, clin. proof-of-concept for structure-based vaccine design may first be achieved for respiratory syncytial virus (RSV), where conformation-dependent access to neutralization-sensitive epitopes on the fusion glycoprotein dets. the capacity to induce potent neutralizing activity. Success with RSV has motivated structure-based stabilization of other class I viral fusion proteins for use as immunogens and demonstrated the importance of structural information for developing vaccines against other viral pathogens, particularly difficult targets that have resisted prior vaccine development efforts. Solving viral surface protein structures also supports rapid vaccine antigen design and application of platform manufg. approaches for emerging pathogens.
- 58Liang, J.; Edelsbrunner, H.; Woodward, C. Anatomy of protein pockets and cavities: measurement of binding site geometry and implications for ligand design. Protein Sci. 1998, 7 (9), 1884– 1897, DOI: 10.1002/pro.556007090558https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK1cXmtFGjsLo%253D&md5=a6f62e03de03da0a6e51361de7a08765Anatomy of protein pockets and cavities: measurement of binding site geometry and implications for ligand designLiang, Jie; Edelsbrunner, Herbert; Woodward, ClareProtein Science (1998), 7 (9), 1884-1897CODEN: PRCIEI; ISSN:0961-8368. (Cambridge University Press)Identification and size characterization of surface pockets and occluded cavities are initial steps in protein structure-based ligand design. A new program, CAST, for automatically locating and measuring protein pockets and cavities, is based on precise computational geometry methods, including alpha shape and discrete flow theory. CAST identifies and measures pockets and pocket mouth openings, as well as cavities. The program specifies the atoms lining pockets, pocket openings, and buried cavities; the vol. and area of pockets and cavities; and the area and circumference of mouth openings. CAST anal. of over 100 proteins has been carried out; proteins examd. include a set of 51 monomeric enzyme-ligand structures, several elastase-inhibitor complexes, the FK506 binding protein, 30 HIV-1 protease-inhibitor complexes, and a no. of small and large protein inhibitors. Medium-sized globular proteins typically have 10-20 pockets/cavities. Most often, binding sites are pockets with 1-2 mouth openings; much less frequently they are cavities. Ligand binding pockets vary widely in size, most within the range 102-103 Å3. Statistical anal. reveals that the no. of pockets and cavities is correlated with protein size, but there is no correlation between the size of the protein and the size of binding sites. Most frequently, the largest pocket/cavity is the active site, but there are a no. of instructive exceptions. Ligand vol. and binding site vol. are somewhat correlated when binding site vol. is ≤700 Å3, but the ligand seldom occupies the entire site. Auxiliary pockets near the active site have been suggested as addnl. binding surface for designed ligands (Mattos C et al., 1994, Nat Struct Biol 1:55-58). Anal. of elastase-inhibitor complexes suggests that CAST can identify ancillary pockets suitable for recruitment in ligand design strategies. Anal. of the FK506 binding protein, and of compds. developed in SAR by NMR (Shuker SB et al., 1996, Science 274:1531-1534), indicates that CAST pocket computation may provide a priori identification of target proteins for linked-fragment design. CAST anal. of 30 HIV-1 protease-inhibitor complexes shows that the flexible active site pocket can vary over a range of 853-1,566 Å3, and that there are two pockets near or adjoining the active site that may be recruited for ligand design.
- 59Naceri, S.; Marc, D.; Blot, R.; Flatters, D.; Camproux, A. C. Druggable Pockets at the RNA Interface Region of Influenza A Virus NS1 Protein Are Conserved across Sequence Variants from Distinct Subtypes. Biomolecules 2023, 13 (1), 64, DOI: 10.3390/biom13010064There is no corresponding record for this reference.
- 60Gavenonis, J.; Sheneman, B. A.; Siegert, T. R.; Eshelman, M. R.; Kritzer, J. A. Comprehensive analysis of loops at protein-protein interfaces for macrocycle design. Nat. Chem. Biol. 2014, 10 (9), 716– 722, DOI: 10.1038/nchembio.158060https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXhtFygs7jE&md5=bf07bd724f173ea74467ab7cd5aedb6aComprehensive analysis of loops at protein-protein interfaces for macrocycle designGavenonis, Jason; Sheneman, Bradley A.; Siegert, Timothy R.; Eshelman, Matthew R.; Kritzer, Joshua A.Nature Chemical Biology (2014), 10 (9), 716-722CODEN: NCBABT; ISSN:1552-4450. (Nature Publishing Group)Inhibiting protein-protein interactions (PPIs) with synthetic mols. remains a frontier of chem. biol. Many PPIs have been successfully targeted by mimicking α-helixes at interfaces, but most PPIs are mediated by nonhelical, nonstrand peptide loops. We sought to comprehensively identify and analyze these loop-mediated PPIs by writing and implementing LoopFinder, a customizable program that can identify loop-mediated PPIs within all of the protein-protein complexes in the Protein Data Bank. Comprehensive anal. of the entire set of 25,005 interface loops revealed common structural motifs and unique features that distinguish loop-mediated PPIs from other PPIs. 'Hot loops', named in analogy to protein hot spots, were identified as loops with favorable properties for mimicry using synthetic mols. The hot loops and their binding partners represent new and promising PPIs for the development of macrocycle and constrained peptide inhibitors.
- 61Corbi-Verge, C.; Kim, P. M. Motif mediated protein-protein interactions as drug targets. Cell Commun. Signal 2016, 14, 8 DOI: 10.1186/s12964-016-0131-4There is no corresponding record for this reference.
- 62Marković, V.; Szczepańska, A.; Berlicki, Ł. Antiviral Protein-Protein Interaction Inhibitors. J. Med. Chem. 2024, 67 (5), 3205– 3231, DOI: 10.1021/acs.jmedchem.3c01543There is no corresponding record for this reference.
- 63Zheng, 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 (4), 331– 332, DOI: 10.1038/nmeth.419363https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXjt1ags7g%253D&md5=5f4e225ef8123dacd8475d526175e1d2MotionCor2: anisotropic correction of beam-induced motion for improved cryo-electron microscopyZheng, Shawn Q.; Palovcak, Eugene; Armache, Jean-Paul; Verba, Kliment A.; Cheng, Yifan; Agard, David A.Nature Methods (2017), 14 (4), 331-332CODEN: NMAEA3; ISSN:1548-7091. (Nature Publishing Group)A review on anisotropic correction of beam-induced motion for improved cryo-electron microscopy. Here we describe MotionCor2, a software tool for anisotropic correction of beam-induced motion. Overall, MotionCor2 is extremely robust and sufficiently accurate at correcting local motions so that the very time-consuming and computationally intensive particle polishing in RELION can be skipped, importantly, it also works on a wide range of data sets, including cryo tomog. tilt series.
- 64Rohou, A.; Grigorieff, N. CTFFIND4: Fast and accurate defocus estimation from electron micrographs. J. Struct. Biol. 2015, 192 (2), 216– 221, DOI: 10.1016/j.jsb.2015.08.00864https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A280%3ADC%252BC287js1Whsg%253D%253D&md5=8500953ad4898ae82de6f8cdc95832cfCTFFIND4: Fast and accurate defocus estimation from electron micrographsRohou Alexis; Grigorieff NikolausJournal of structural biology (2015), 192 (2), 216-21 ISSN:.CTFFIND is a widely-used program for the estimation of objective lens defocus parameters from transmission electron micrographs. Defocus parameters are estimated by fitting a model of the microscope's contrast transfer function (CTF) to an image's amplitude spectrum. Here we describe modifications to the algorithm which make it significantly faster and more suitable for use with images collected using modern technologies such as dose fractionation and phase plates. We show that this new version preserves the accuracy of the original algorithm while allowing for higher throughput. We also describe a measure of the quality of the fit as a function of spatial frequency and suggest this can be used to define the highest resolution at which CTF oscillations were successfully modeled.
- 65Punjani, A.; Rubinstein, J. L.; Fleet, D. J.; Brubaker, M. A. cryoSPARC: algorithms for rapid unsupervised cryo-EM structure determination. Nat. Methods 2017, 14 (3), 290– 296, DOI: 10.1038/nmeth.416965https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXitlGisbs%253D&md5=95d468147707707e70ac0ad38dd6ebf6cryoSPARC: algorithms for rapid unsupervised cryo-EM structure determinationPunjani, Ali; Rubinstein, John L.; Fleet, David J.; Brubaker, Marcus A.Nature Methods (2017), 14 (3), 290-296CODEN: NMAEA3; ISSN:1548-7091. (Nature Publishing Group)Single-particle electron cryomicroscopy (cryo-EM) is a powerful method for detg. the structures of biol. macromols. With automated microscopes, cryo-EM data can often be obtained in a few days. However, processing cryo-EM image data to reveal heterogeneity in the protein structure and to refine 3D maps to high resoln. frequently becomes a severe bottleneck, requiring expert intervention, prior structural knowledge, and weeks of calcns. on expensive computer clusters. Here we show that stochastic gradient descent (SGD) and branch-and-bound max. likelihood optimization algorithms permit the major steps in cryo-EM structure detn. to be performed in hours or minutes on an inexpensive desktop computer. Furthermore, SGD with Bayesian marginalization allows ab initio 3D classification, enabling automated anal. and discovery of unexpected structures without bias from a ref. map. These algorithms are combined in a user-friendly computer program named cryoSPARC (http://www.cryosparc.com).
- 66Yang, Z.; Lasker, K.; Schneidman-Duhovny, D.; Webb, B.; Huang, C. C.; Pettersen, E. F.; Goddard, T. D.; Meng, E. C.; Sali, A.; Ferrin, T. E. UCSF Chimera, MODELLER, and IMP: an integrated modeling system. J. Struct Biol. 2012, 179 (3), 269– 278, DOI: 10.1016/j.jsb.2011.09.006There is no corresponding record for this reference.
- 67Dam, J.; Schuck, P. Calculating Sedimentation Coefficient Distributions by Direct Modeling of Sedimentation Velocity Concentration Profiles. In Methods in Enzymology; Academic Press, 2004; Vol. 384, pp 185– 212.There is no corresponding record for this reference.
- 68Ribaric, S.; Peterec, D.; Sketelj, J. Computer aided data acquisition and analysis of acetlycholinesterase velocity sedimentation profiles. Comput. Methods Program. Biomed. 1996, 49 (2), 149– 156, DOI: 10.1016/0169-2607(96)01719-1There is no corresponding record for this reference.
- 69Johnson, B. A.; Blevins, R. A. NMR View: A computer program for the visualization and analysis of NMR data. J. Biomol. NMR 1994, 4 (5), 603– 614, DOI: 10.1007/BF0040427269https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2cXmt1Gkurw%253D&md5=6a93a0d9704fbbb654abc192203c68aeNMRView: a computer program for the visualization and analysis of NMR dataJohnson, Bruce A.; Blevins, Richard A.Journal of Biomolecular NMR (1994), 4 (5), 603-14CODEN: JBNME9; ISSN:0925-2738.NMRView is a computer program designed for the visualization and anal. of NMR data. It allows the user to interact with a practically unlimited no. of 2D, 3D and 4D NMR data files. Any no. of spectral windows can be displayed on the screen in any size and location. Automatic peak picking and facilitated peak anal. features are included to aid in the assignment of complex NMR spectra. NMR View provides structure anal. features and data transfer to and from structural generation programs, allowing for a tight coupling between the spectral anal. and structural generation.,. Visual correlation between structures and spectra can be done with the Mol. Data Viewer, a mol. graphics program with bidirectional communication to NMR View. The used interface can be customized and a command language is provided to allow for the automation of various tasks.
- 70Schwieters, C. D.; Kuszewski, J. J.; Marius Clore, G. Using Xplor–NIH for NMR molecular structure determination. Prog. Nucl. Magn. Reson. Spectrosc. 2006, 48 (1), 47– 62, DOI: 10.1016/j.pnmrs.2005.10.00170https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD28XjtVOiurY%253D&md5=56257dc51ddafdf8c467d30aef191ee2Using Xplor-NIH for NMR molecular structure determinationSchwieters, Charles D.; Kuszewski, John J.; Clore, G. MariusProgress in Nuclear Magnetic Resonance Spectroscopy (2006), 48 (1), 47-62CODEN: PNMRAT; ISSN:0079-6565. (Elsevier B.V.)A review of the title program for computerized structure detn.
- 71Maciejewski, M. W.; Schuyler, A. D.; Gryk, M. R.; Moraru, I. I.; Romero, P. R.; Ulrich, E. L.; Eghbalnia, H. R.; Livny, M.; Delaglio, F.; Hoch, J. C. NMRbox: A Resource for Biomolecular NMR Computation. Biophys. J. 2017, 112 (8), 1529– 1534, DOI: 10.1016/j.bpj.2017.03.01171https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXlvFWgtL8%253D&md5=a990d67aed35b62cd010f0f5396d0f82NMRbox: A Resource for Biomolecular NMR ComputationMaciejewski, Mark W.; Schuyler, Adam D.; Gryk, Michael R.; Moraru, Ion I.; Romero, Pedro R.; Ulrich, Eldon L.; Eghbalnia, Hamid R.; Livny, Miron; Delaglio, Frank; Hoch, Jeffrey C.Biophysical Journal (2017), 112 (8), 1529-1534CODEN: BIOJAU; ISSN:0006-3495. (Cell Press)Advances in computation have been enabling many recent advances in biomol. applications of NMR. Due to the wide diversity of applications of NMR, the no. and variety of software packages for processing and analyzing NMR data is quite large, with labs relying on dozens, if not hundreds of software packages. Discovery, acquisition, installation, and maintenance of all these packages is a burdensome task. Because the majority of software packages originate in academic labs, persistence of the software is compromised when developers graduate, funding ceases, or investigators turn to other projects. To simplify access to and use of biomol. NMR software, foster persistence, and enhance reproducibility of computational workflows, the authors have developed NMRbox, a shared resource for NMR software and computation. NMRbox employs virtualization to provide a comprehensive software environment preconfigured with hundreds of software packages, available as a downloadable virtual machine or as a Platform-as-a-Service supported by a dedicated compute cloud. Ongoing development includes a metadata harvester to regularize, annotate, and preserve workflows and facilitate and enhance data depositions to BioMagResBank, and tools for Bayesian inference to enhance the robustness and extensibility of computational analyses. In addn. to facilitating use and preservation of the rich and dynamic software environment for biomol. NMR, NMRbox fosters the development and deployment of a new class of metasoftware packages. NMRbox is freely available to not-for-profit users.
- 72Shen, Y.; Delaglio, F.; Cornilescu, G.; Bax, A. TALOS+: a hybrid method for predicting protein backbone torsion angles from NMR chemical shifts. J. Biomol. NMR 2009, 44 (4), 213– 223, DOI: 10.1007/s10858-009-9333-z72https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXovFajtr8%253D&md5=ee3f8647a02e19937988056fd7af04b2TALOS+: a hybrid method for predicting protein backbone torsion angles from NMR chemical shiftsShen, Yang; Delaglio, Frank; Cornilescu, Gabriel; Bax, AdJournal of Biomolecular NMR (2009), 44 (4), 213-223CODEN: JBNME9; ISSN:0925-2738. (Springer)NMR chem. shifts in proteins depend strongly on local structure. The program TALOS establishes an empirical relation between 13C, 15N and 1H chem. shifts and backbone torsion angles .vphi. and ψ. Extension of the original 20-protein database to 200 proteins increased the fraction of residues for which backbone angles could be predicted from 65 to 74%, while reducing the error rate from 3 to 2.5%. Addn. of a two-layer neural network filter to the database fragment selection process forms the basis for a new program, TALOS+, which further enhances the prediction rate to 88.5%, without increasing the error rate. Excluding the 2.5% of residues for which TALOS+ makes predictions that strongly differ from those obsd. in the cryst. state, the accuracy of predicted .vphi. and ψ angles, equals ±13°. Large discrepancies between predictions and crystal structures are primarily limited to loop regions, and for the few cases where multiple X-ray structures are available such residues are often found in different states in the different structures. The TALOS+ output includes predictions for individual residues with missing chem. shifts, and the neural network component of the program also predicts secondary structure with good accuracy.
- 73Tian, Y.; Schwieters, C. D.; Opella, S. J.; Marassi, F. M. A practical implicit solvent potential for NMR structure calculation. J. Magn. Reson. 2014, 243, 54– 64, DOI: 10.1016/j.jmr.2014.03.01173https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2cXovVahu70%253D&md5=68ad01c9ceb6078a4291b0d6e96408e8A practical implicit solvent potential for NMR structure calculationTian, Ye; Schwieters, Charles D.; Opella, Stanley J.; Marassi, Francesca M.Journal of Magnetic Resonance (2014), 243 (), 54-64CODEN: JMARF3; ISSN:1090-7807. (Elsevier B.V.)The benefits of protein structure refinement in water are well documented. However, performing structure refinement with explicit at. representation of the solvent mols. is computationally expensive and impractical for NMR-restrained structure calcns. that start from completely extended polypeptide templates. Here we describe a new implicit solvation potential, EEFx (Effective Energy Function for XPLOR-NIH), for NMR-restrained structure calcns. of proteins in XPLOR-NIH. The key components of EEFx are an energy term for solvation energy that works together with other nonbonded energy functions, and a dedicated force field for conformational and nonbonded protein interaction parameters. The initial results obtained with EEFx show that significant improvements in structural quality can be obtained. EEFx is computationally efficient and can be used both to fold and refine structures. Overall, EEFx improves the quality of protein conformation and nonbonded at. interactions. Moreover, such benefits are accompanied by enhanced structural precision and enhanced structural accuracy, reflected in improved agreement with the cross-validated dipolar coupling data. Finally, implementation of EEFx calcns. is straightforward and computationally efficient. Overall, EEFx provides a useful method for the practical calcn. of exptl. protein structures in a phys. realistic environment.
- 74Gore, S.; Sanz García, E.; Hendrickx, P. M. S.; Gutmanas, A.; Westbrook, J. D.; Yang, H.; Feng, Z.; Baskaran, K.; Berrisford, J. M.; Hudson, B. P. Validation of Structures in the Protein Data Bank. Structure 2017, 25 (12), 1916– 1927, DOI: 10.1016/j.str.2017.10.00974https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXhvVOitbnE&md5=680c220ca1354e8a089638ae978501d3Validation of Structures in the Protein Data BankGore, Swanand; Sanz Garcia, Eduardo; Hendrickx, Pieter M. S.; Gutmanas, Aleksandras; Westbrook, John D.; Yang, Huanwang; Feng, Zukang; Baskaran, Kumaran; Berrisford, John M.; Hudson, Brian P.; Ikegawa, Yasuyo; Kobayashi, Naohiro; Lawson, Catherine L.; Mading, Steve; Mak, Lora; Mukhopadhyay, Abhik; Oldfield, Thomas J.; Patwardhan, Ardan; Peisach, Ezra; Sahni, Gaurav; Sekharan, Monica R.; Sen, Sanchayita; Shao, Chenghua; Smart, Oliver S.; Ulrich, Eldon L.; Yamashita, Reiko; Quesada, Martha; Young, Jasmine Y.; Nakamura, Haruki; Markley, John L.; Berman, Helen M.; Burley, Stephen K.; Velankar, Sameer; Kleywegt, Gerard J.Structure (Oxford, United Kingdom) (2017), 25 (12), 1916-1927CODEN: STRUE6; ISSN:0969-2126. (Elsevier Ltd.)The Worldwide PDB recently launched a deposition, biocuration, and validation tool: OneDep. At various stages of OneDep data processing, validation reports for three-dimensional structures of biol. macromols. are produced. These reports are based on recommendations of expert task forces representing crystallog., NMR, and cryoelectron microscopy communities. The reports provide useful metrics with which depositors can evaluate the quality of the exptl. data, the structural model, and the fit between them. The validation module is also available as a stand-alone web server and as a programmatically accessible web service. A growing no. of journals require the official wwPDB validation reports (produced at biocuration) to accompany manuscripts describing macromol. structures. Upon public release of the structure, the validation report becomes part of the public PDB archive. Geometric quality scores for proteins in the PDB archive have improved over the past decade.
- 75Schwarzinger, S.; Kroon, G. J.; Foss, T. R.; Wright, P. E.; Dyson, H. J. Random coil chemical shifts in acidic 8 M urea: implementation of random coil shift data in NMRView. J. Biomol. NMR 2000, 18 (1), 43– 48, DOI: 10.1023/A:100838681652175https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3cXnvVGit7w%253D&md5=cf30cd19aa4420c807a0525789db9be1Random coil chemical shifts in acidic 8 M urea: implementation of random coil shift data in NMRViewSchwarzinger, Stephan; Kroon, Gerard J. A.; Foss, Ted R.; Wright, Peter E.; Dyson, H. JaneJournal of Biomolecular NMR (2000), 18 (1), 43-48CODEN: JBNME9; ISSN:0925-2738. (Kluwer Academic Publishers)Studies of proteins unfolded in acid or chem. denaturant can help in unraveling events during the earliest phases of protein folding. In order for meaningful comparisons to be made of residual structure in unfolded states, it is necessary to use random coil chem. shifts that are valid for the exptl. system under study. We present a set of random coil chem. shifts obtained for model peptides under exptl. conditions used in studies of denatured proteins. This new set, together with previously published data sets, has been incorporated into a software interface for NMRView, allowing selection of the random coil data set that fits the exptl. conditions best.
- 76Berendsen, H. J. C.; van der Spoel, D.; van Drunen, R. GROMACS: A message-passing parallel molecular dynamics implementation. Comput. Phys. Commun. 1995, 91 (1), 43– 56, DOI: 10.1016/0010-4655(95)00042-E76https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXps1Wrtr0%253D&md5=04d823aeab28ca374efb86839c705179GROMACS: A message-passing parallel molecular dynamics implementationBerendsen, H. J. C.; van der Spoel, D.; van Drunen, R.Computer Physics Communications (1995), 91 (1-3), 43-56CODEN: CPHCBZ; ISSN:0010-4655. (Elsevier)A parallel message-passing implementation of a mol. dynamics (MD) program that is useful for bio(macro)mols. in aq. environment is described. The software has been developed for a custom-designed 32-processor ring GROMACS (Groningen MAchine for Chem. Simulation) with communication to and from left and right neighbors, but can run on any parallel system onto which a a ring of processors can be mapped and which supports PVM-like block send and receive calls. The GROMACS software consists of a preprocessor, a parallel MD and energy minimization program that can use an arbitrary no. of processors (including one), an optional monitor, and several anal. tools. The programs are written in ANSI C and available by ftp (information: [email protected]). The functionality is based on the GROMOS (Groningen Mol. Simulation) package (van Gunsteren and Berendsen, 1987; BIOMOS B.V., Nijenborgh 4, 9747 AG Groningen). Conversion programs between GROMOS and GROMACS formats are included.The MD program can handle rectangular periodic boundary conditions with temp. and pressure scaling. The interactions that can be handled without modification are variable non-bonded pair interactions with Coulomb and Lennard-Jones or Buckingham potentials, using a twin-range cut-off based on charge groups, and fixed bonded interactions of either harmonic or constraint type for bonds and bond angles and either periodic or cosine power series interactions for dihedral angles. Special forces can be added to groups of particles (for non-equil. dynamics or for position restraining) or between particles (for distance restraints). The parallelism is based on particle decompn. Interprocessor communication is largely limited to position and force distribution over the ring once per time step.
- 77Jurrus, E.; Engel, D.; Star, K.; Monson, K.; Brandi, J.; Felberg, L. E.; Brookes, D. H.; Wilson, L.; Chen, J.; Liles, K. Improvements to the APBS biomolecular solvation software suite. Protein Sci. 2018, 27 (1), 112– 128, DOI: 10.1002/pro.328077https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC2sXhslSkt7vI&md5=99651d125e38f4a85d453fecf0f71652Improvements to the APBS biomolecular solvation software suiteJurrus, Elizabeth; Engel, Dave; Star, Keith; Monson, Kyle; Brandi, Juan; Felberg, Lisa E.; Brookes, David H.; Wilson, Leighton; Chen, Jiahui; Liles, Karina; Chun, Minju; Li, Peter; Gohara, David W.; Dolinsky, Todd; Konecny, Robert; Koes, David R.; Nielsen, Jens Erik; Head-Gordon, Teresa; Geng, Weihua; Krasny, Robert; Wei, Guo-Wei; Holst, Michael J.; McCammon, J. Andrew; Baker, Nathan A.Protein Science (2018), 27 (1), 112-128CODEN: PRCIEI; ISSN:1469-896X. (Wiley-Blackwell)The Adaptive Poisson-Boltzmann Solver (APBS) software was developed to solve the equations of continuum electrostatics for large biomol. assemblages that have provided impact in the study of a broad range of chem., biol., and biomedical applications. APBS addresses the three key technol. challenges for understanding solvation and electrostatics in biomedical applications: accurate and efficient models for biomol. solvation and electrostatics, robust and scalable software for applying those theories to biomol. systems, and mechanisms for sharing and analyzing biomol. electrostatics data in the scientific community. To address new research applications and advancing computational capabilities, we have continually updated APBS and its suite of accompanying software since its release in 2001. In this article, we discuss the models and capabilities that have recently been implemented within the APBS software package including a Poisson-Boltzmann anal. and a semi-anal. solver, an optimized boundary element solver, a geometry-based geometric flow solvation model, a graph theory-based algorithm for detg. pKa values, and an improved web-based visualization tool for viewing electrostatics.
- 78Martínez, L.; Andrade, R.; Birgin, E. G.; Martínez, J. M. PACKMOL: A package for building initial configurations for molecular dynamics simulations. J. Comput. Chem. 2009, 30 (13), 2157– 2164, DOI: 10.1002/jcc.2122478https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD1MXptleqsb8%253D&md5=2a76255c873b866a26540f7e84496272PACKMOL: A package for building initial configurations for molecular dynamics simulationsMartinez, L.; Andrade, R.; Birgin, E. G.; Martinez, J. M.Journal of Computational Chemistry (2009), 30 (13), 2157-2164CODEN: JCCHDD; ISSN:0192-8651. (John Wiley & Sons, Inc.)Adequate initial configurations for mol. dynamics simulations consist of arrangements of mols. distributed in space in such a way to approx. represent the system's overall structure. In order that the simulations are not disrupted by large van der Waals repulsive interactions, atoms from different mols. must keep safe pairwise distances. Obtaining such a mol. arrangement can be considered a packing problem: Each type mol. must satisfy spatial constraints related to the geometry of the system, and the distance between atoms of different mols. must be greater than some specified tolerance. We have developed a code able to pack millions of atoms, grouped in arbitrarily complex mols., inside a variety of three-dimensional regions. The regions may be intersections of spheres, ellipses, cylinders, planes, or boxes. The user must provide only the structure of one mol. of each type and the geometrical constraints that each type of mol. must satisfy. Building complex mixts., interfaces, solvating biomols. in water, other solvents, or mixts. of solvents, is straightforward. In addn., different atoms belonging to the same mol. may also be restricted to different spatial regions, in such a way that more ordered mol. arrangements can be built, as micelles, lipid double-layers, etc. The packing time for state-of-the-art mol. dynamics systems varies from a few seconds to a few minutes in a personal computer. The input files are simple and currently compatible with PDB, Tinker, Molden, or Moldy coordinate files. The package is distributed as free software and can be downloaded from . © 2009 Wiley Periodicals, Inc. J Comput Chem, 2009.
- 79Rühle, V. Pressure coupling/barostats. 2008.There is no corresponding record for this reference.
- 80Essmann, U.; Perera, L.; Berkowitz, M. L.; Darden, T.; Lee, H.; Pedersen, L. G. A smooth particle mesh Ewald method. J. Chem. Phys. 1995, 103 (19), 8577– 8593, DOI: 10.1063/1.47011780https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaK2MXptlehtrw%253D&md5=092a679dd3bee08da28df41e302383a7A smooth particle mesh Ewald methodEssmann, Ulrich; Perera, Lalith; Berkowitz, Max L.; Darden, Tom; Lee, Hsing; Pedersen, Lee G.Journal of Chemical Physics (1995), 103 (19), 8577-93CODEN: JCPSA6; ISSN:0021-9606. (American Institute of Physics)The previously developed particle mesh Ewald method is reformulated in terms of efficient B-spline interpolation of the structure factors. This reformulation allows a natural extension of the method to potentials of the form 1/rp with p ≥ 1. Furthermore, efficient calcn. of the virial tensor follows. Use of B-splines in the place of Lagrange interpolation leads to analytic gradients as well as a significant improvement in the accuracy. The authors demonstrate that arbitrary accuracy can be achieved, independent of system size N, at a cost that scales as N log(N). For biomol. systems with many thousands of atoms and this method permits the use of Ewald summation at a computational cost comparable to that of a simple truncation method of 10 Å or less.
- 81Lindorff-Larsen, K.; Piana, S.; Palmo, K.; Maragakis, P.; Klepeis, J. L.; Dror, R. O.; Shaw, D. E. Improved side-chain torsion potentials for the Amber ff99SB protein force field. Proteins 2010, 78, 1950– 1958, DOI: 10.1002/prot.2271181https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXkvFegtLo%253D&md5=447a9004026e2b93f0f7beff165daa09Improved side-chain torsion potentials for the Amber ff99SB protein force fieldLindorff-Larsen, Kresten; Piana, Stefano; Palmo, Kim; Maragakis, Paul; Klepeis, John L.; Dror, Ron O.; Shaw, David E.Proteins: Structure, Function, and Bioinformatics (2010), 78 (8), 1950-1958CODEN: PSFBAF ISSN:. (Wiley-Liss, Inc.)Recent advances in hardware and software have enabled increasingly long mol. dynamics (MD) simulations of biomols., exposing certain limitations in the accuracy of the force fields used for such simulations and spurring efforts to refine these force fields. Recent modifications to the Amber and CHARMM protein force fields, for example, have improved the backbone torsion potentials, remedying deficiencies in earlier versions. Here, the authors further advance simulation accuracy by improving the amino acid side-chain torsion potentials of the Amber ff99SB force field. First, the authors used simulations of model alpha-helical systems to identify the four residue types whose rotamer distribution differed the most from expectations based on Protein Data Bank statistics. Second, the authors optimized the side-chain torsion potentials of these residues to match new, high-level quantum-mech. calcns. Finally, the authors used microsecond-timescale MD simulations in explicit solvent to validate the resulting force field against a large set of exptl. NMR measurements that directly probe side-chain conformations. The new force field, which the authors have termed Amber ff99SB-ILDN, exhibits considerably better agreement with the NMR data. Proteins 2010. © 2010 Wiley-Liss, Inc.
- 82Mark, P.; Nilsson, L. Structure and Dynamics of the TIP3P, SPC, and SPC/E Water Models at 298 K. J. Phys. Chem. A 2001, 105 (43), 9954– 9960, DOI: 10.1021/jp003020w82https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BD3MXntlWrurs%253D&md5=fecd3db40210b04e8b2a933ea07b131eStructure and Dynamics of the TIP3P, SPC, and SPC/E Water Models at 298 KMark, Pekka; Nilsson, LennartJournal of Physical Chemistry A (2001), 105 (43), 9954-9960CODEN: JPCAFH; ISSN:1089-5639. (American Chemical Society)Mol. dynamics simulations of five water models, the TIP3P (original and modified), SPC (original and refined), and SPC/E (original), were performed using the CHARMM mol. mechanics program. All simulations were carried out in the microcanonical NVE ensemble, using 901 water mols. in a cubic simulation cell furnished with periodic boundary conditions at 298 K. The SHAKE algorithm was used to keep water mols. rigid. Nanosecond trajectories were calcd. with all water models for high statistical accuracy. The characteristic self-diffusion coeffs. D and radial distribution functions, gOO, gOH, and gHH for all five water models were detd. and compared to exptl. data. The effects of velocity rescaling on the self-diffusion coeff. D were examd. All these empirical water models used in this study are similar by having three interaction sites, but the small differences in their pair potentials composed of Lennard-Jones (LJ) and Coulombic terms give significant differences in the calcd. self-diffusion coeffs., and in the height of the second peak of the radial distribution function gOO.
- 83de Vries, S. J.; van Dijk, M.; Bonvin, A. M. J. J. The HADDOCK web server for data-driven biomolecular docking. Nat. Protoc. 2010, 5 (5), 883– 897, DOI: 10.1038/nprot.2010.3283https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADC%252BC3cXltl2qsLw%253D&md5=dbf801d99065e8d2c3aedc016209838cThe HADDOCK web server for data-driven biomolecular dockingde Vries, Sjoerd J.; van Dijk, Marc; Bonvin, Alexandre M. J. J.Nature Protocols (2010), 5 (5), 883-897CODEN: NPARDW; ISSN:1750-2799. (Nature Publishing Group)Computational docking is the prediction or modeling of the three-dimensional structure of a biomol. complex, starting from the structures of the individual mols. in their free, unbound form. HADDOCK is a popular docking program that takes a data-driven approach to docking, with support for a wide range of exptl. data. Here the authors present the HADDOCK web server protocol, facilitating the modeling of biomol. complexes for a wide community. The main web interface is user-friendly, requiring only the structures of the individual components and a list of interacting residues as input. Addnl. web interfaces allow the more advanced user to exploit the full range of exptl. data supported by HADDOCK and to customize the docking process. The HADDOCK server has access to the resources of a dedicated cluster and of the e-NMR GRID infrastructure. Therefore, a typical docking run takes only a few minutes to prep. and a few hours to complete.
- 84Xue, L. C.; Rodrigues, J. P.; Kastritis, P. L.; Bonvin, A. M.; Vangone, A. PRODIGY: a web server for predicting the binding affinity of protein-protein complexes. Bioinformatics 2016, 32 (23), 3676– 3678, DOI: 10.1093/bioinformatics/btw514There is no corresponding record for this reference.
- 85Schuck, P.; Zhao, H. Sedimentation Velocity Analytical Ultracentrifugation: Interacting Systems (1st ed.); CRC Press, 2017.There is no corresponding record for this reference.
- 86Kyte, J.; Doolittle, R. F. A simple method for displaying the hydropathic character of a protein. J. Mol. Biol. 1982, 157 (1), 105– 132, DOI: 10.1016/0022-2836(82)90515-086https://chemport.cas.org/services/resolver?origin=ACS&resolution=options&coi=1%3ACAS%3A528%3ADyaL38Xks1yjtro%253D&md5=ee67eb115939dfe56b2b2cae2c32dbd3A simple method for displaying the hydropathic character of a proteinKyte, Jack; Doolittle, Russell F.Journal of Molecular Biology (1982), 157 (1), 105-32CODEN: JMOBAK; ISSN:0022-2836.A computer program that progressively evaluates the hydrophilicity and hydrophobicity of a protein along its amino acid sequence was devised. A hydropathy scale takes into consideration the hydrophilic and hydrophobic properties of each of the 20 amino acid side chains. Correlation was demonstrated between the plotted values and known structures detd. by crystallog.
Supporting Information
Supporting Information
The Supporting Information is available free of charge at https://pubs.acs.org/doi/10.1021/acsinfecdis.4c00407.
Experimental methods; Figures S1–S21; Tables S1 and S2 (PDF)
Conformational change of GNNV VLP from pH 8.0 to 5.0 (Movie S1) (MP4)
Enlarged views of GNNV protrusion conformational change from pH 8.0 to 5.0 (Movie S2) (MP4)
Hypothetical atomic model of GNNV-P conformational change from pH 8.0 to 5.0 (Movie S3) (MP4)
Terms & Conditions
Most electronic Supporting Information files are available without a subscription to ACS Web Editions. Such files may be downloaded by article for research use (if there is a public use license linked to the relevant article, that license may permit other uses). Permission may be obtained from ACS for other uses through requests via the RightsLink permission system: http://pubs.acs.org/page/copyright/permissions.html.