Conformation and Lipid Interaction of the Fusion Peptide of the Paramyxovirus PIV5 in Anionic and Negative-Curvature Membranes from Solid-State NMRClick to copy article linkArticle link copied!
Abstract
Viral fusion proteins catalyze the merger of the virus envelope and the target cell membrane through multiple steps of protein conformational changes. The fusion peptide domain of these proteins is important for membrane fusion, but how it causes membrane curvature and dehydration is still poorly understood. We now use solid-state NMR spectroscopy to investigate the conformation, topology, and lipid and water interactions of the fusion peptide of the PIV5 virus F protein in three lipid membranes, POPC/POPG, DOPC/DOPG, and DOPE. These membranes allow us to investigate the effects of lipid chain disorder, membrane surface charge, and intrinsic negative curvature on the fusion peptide structure. Chemical shifts and spin diffusion data indicate that the PIV5 fusion peptide is inserted into all three membranes but adopts distinct conformations: it is fully α-helical in the POPC/POPG membrane, adopts a mixed strand/helix conformation in the DOPC/DOPG membrane, and is primarily a β-strand in the DOPE membrane. 31P NMR spectra show that the peptide retains the lamellar structure and hydration of the two anionic membranes. However, it dehydrates the DOPE membrane, destabilizes its inverted hexagonal phase, and creates an isotropic phase that is most likely a cubic phase. The ability of the β-strand conformation of the fusion peptide to generate negative Gaussian curvature and to dehydrate the membrane may be important for the formation of hemifusion intermediates in the membrane fusion pathway.
Introduction
Experimental Section
Peptide and Lipids
sample | labeled residue |
---|---|
GVAL | G114, V115, A126, L127 |
IGALV | I108, G109, A112, L113, V125 |
GVTAA | G105, V106, T122, A123, A124 |
VLAAT | V107, L110, A111, A116, T117 |
AAQV | A118, A119, Q120, V121 |
FPK4(103–129): FAGVVIGLAALGVATAAQVTAAVALVK-DIOXA-KKKK.
Solid-State NMR Experiments
Results
Complete Backbone Conformation of FPK4 in the POPC/POPG Membrane
Figure 1
Figure 1. 2D 13C–13C correlation spectra of PIV5 FPK4 in gel-phase POPC/POPG (4:1) bilayers. Shown at the top is the amino acid sequence with labeled residues color-coded according to samples. (a) GVTAA-FPK4 spectrum, measured at 253 K with 20 ms mixing. (b) VLAAT-FPK4 spectrum, coadded from two spectra measured at 243 K with 20 ms mixing and 253 K with 60 ms mixing. (c) AAQV-FPK4 spectrum, measured at 253 K with 20 ms mixing. All residues show resolved and α-helical chemical shifts. Superscript h denotes helical chemical shifts.
Figure 2
Figure 2. 2D 15N–13C correlation spectra of PIV5 FPK4 in gel-phase POPC/POPG (magenta), DOPC/DOPG (black), and POPC (blue) membranes. (a) GVTAA-FPK4 spectra. (b) VLAAT-FPK4 spectra. (c) IGALV-FPK4 spectra. (d) GVAL-FPK4 spectra. (e) AAQV-FPK4 spectra. The peptide shows predominantly β-strand chemical shifts in the POPC membrane, α-helical chemical shifts in the POPC/POPG membrane, and mixed strand and helix chemical shifts in the DOPC/DOPG membrane. Most residues in the GVTAA and VLAAT samples show two sets of chemical shifts in the DOPC/DOPG bilayer. The AAQV sample shows nearly identical α-helical chemical shifts in the POPC/POPG and DOPC/DOPG membranes. Superscripts h and s denote helical and strand chemical shifts, respectively.
POPC/POPG (4:1) | DOPC/DOPG (4:1) | |||||||||||
---|---|---|---|---|---|---|---|---|---|---|---|---|
residue | N | CO | Cα | Cβ | Cγ | Cδ | N | CO | Cα | Cβ | Cγ | Cδ |
G105b | 105.2 | 173.7 | 44.8 | 106.6 | 168.4 | 43.7 | ||||||
V106b | 122.8 | 174.8 | 64.9 | 29.6 | 21.9/19.1 | 118.6 | 170.0 | 56.8 | 34.0 | 19.5/18.4 | ||
V107c | 118.2 | 175.2 | 64.9 | 29.6 | 20.2 | 123.4 | 171.8 | 57.2 | 32.8 | 19.4 | ||
117.4 | 175.1 | 64.8 | 29.3 | 20.8/19.6 | ||||||||
I108d | 117.7 | 175.6 | 63.3 | 36.3 | 28.6/15.4 | 13.0 | 123.8 | 171.6 | 56.3 | 40.0 | 26.4/15.2 | 12.6 |
G109d | 107.7 | 173.6 | 45.5 | 112.1 | 168.3 | 42.3 | ||||||
L110c | 121.8 | 176.6 | 55.3 | 39.8 | 24.7 | 122.3 | 171.8 | 51.3 | 44.7 | 24.6 | ||
121.4 | 176.7 | 55.5 | 39.5 | 24.6 | 20.7 | |||||||
A111c | 121.4 | 177.0 | 53.1 | 15.5 | 124.1 | 172.5 | 48.6 | 20.8 | ||||
121.4 | 176.9 | 53.0 | 15.9 | |||||||||
A112d | 119.3 | 176.6 | 53.1 | 16.4 | 121.9 | 172.1 | 48.3 | 21.7 | ||||
176.5 | 53.2 | 15.6 | ||||||||||
L113d | 119.9 | 177.9 | 55.2 | 39.8 | 24.7 | 20.0 | 121.1 | 172.2 | 51.1 | 44.5 | 25.0 | 22.0 |
178.2 | 55.2 | 39.3 | 24.6 | |||||||||
G114e | 110.4 | 171.9 | 45.7 | ND | ||||||||
V115e | 122.4 | 175.2 | 64.4 | 29.3 | 20.6/19.2 | ND | ||||||
A116c | 121.4 | 176.7 | 53.1 | 16.6 | 121.5 | 176.9 | 53.0 | 15.9 | ||||
124.5 | 172.5 | 48.6 | 20.8 | |||||||||
T117c | 115.8 | 173.8 | 65.5 | 65.5 | 18.8 | 115.1 | 173.7 | 65.2 | 65.2 | 19.2 | ||
123.8 | 173.3 | 57.6 | 68.5 | 20.0 | ||||||||
A118f | 124.6 | 178.6 | 53.3 | 16.2 | 123.6 | 178.6 | 52.8 | 16.0 | ||||
173.1 | 48.7 | 21.0 | ||||||||||
A119f | 124.6 | 176.6 | 53.3 | 16.3 | 120.4 | 176.6 | 52.8 | 16.0 | ||||
173.1 | 48.7 | 21.0 | ||||||||||
Q120f | 119.6 | 177.5 | 56.1 | 27.2 | 32.0 | 120.4 | 176.8 | 56.4 | 25.7 | 31.9 | ||
V121f | 120.8 | 175.1 | 64.7 | 29.5 | 20.2 | 119.9 | 175.3 | 64.7 | 29.2 | 19.9 | ||
57.8 | 19.5 | |||||||||||
T122b | 115.2 | 173.6 | 65.9 | 65.9 | 19.7 | 116.4 | 174.0 | 65.7 | 65.7 | 20.0 | ||
123.2 | 170.5 | 59.3 | 69.2 | 19.9 | ||||||||
A123b | 121.9 | 175.8 | 53.1 | 17.3 | 122.6 | 177.0 | 52.7 | 16.3 | ||||
121.4 | 172.6 | 48.8 | 21.6 | |||||||||
A124b | 119.1 | 176.5 | 53.2 | 16.2 | 122.6 | 177.0 | 52.7 | 16.3 | ||||
121.4 | 172.6 | 48.8 | 21.6 | |||||||||
V125d | 116.7 | 175.3 | 64.1 | 29.1 | 20.6 | 117.1 | 175.4 | 64.1 | 29.2 | 21.0/19.8 | ||
A126e | 120.0 | 177.3 | 52.6 | 15.9 | ND | |||||||
L127e | 118.6 | 175.7 | 55.0 | 40.0 | 24.8 | 21.5 | ND |
Chemical shifts were measured from 2D spectra at 233–253 K. Italics indicate the second conformation. 13C chemical shifts are referenced to TMS, and 15N chemical shifts are referenced to liquid ammonia.
From the GVTAA sample (G105, V106, T122, A123, and A124).
From the VLAAT sample (V107, L110, A111, A116, and T117).
From the IGALV sample (I108, G109, A112, L113, and V125).
From the GVAL sample (G114, V115, A126, and L127).
From the AAQV sample (A118, A119, Q120, and V121).
Figure 3
Figure 3. 13C and 15N secondary chemical shifts of FPK4 in (a) POPC/POPG and (b) DOPC/DOPG membranes. FPK4 shows clear α-helical chemical shifts (red) in POPC/POPG bilayers and mixed helical and strand chemical shifts (blue) in DOPC/DOPG bilayers. The random coil values of Zhang et al. (86) were used to calculate the secondary shifts.
POPC/POPG (4:1) | DOPC/DOPG (4:1) | |||
---|---|---|---|---|
residue | φ (deg) | ψ (deg) | φ (deg) | ψ (deg) |
V106 | –65 ± 4 | –44 ± 5 | –135 ± 19 | 146 ± 17 |
V107 | –61 ± 6 | –41 ± 10 | –124 ± 11 | 140 ± 17 |
I108 | –64 ± 10 | –38 ± 16 | –130 ± 13 | 143 ± 15 |
G109 | –63 ± 5 | –39 ± 4 | –142 ± 19 | 156 ± 18 |
L110 | –64 ± 4 | –42 ± 7 | –136 ± 11 | 146 ± 12 |
A111 | –65 ± 7 | –39 ± 6 | –131 ± 17 | 147 ± 12 |
A112 | –63 ± 6 | –42 ± 7 | –136 ± 11 | 146 ± 15 |
L113 | –66 ± 6 | –38 ± 7 | –123 ± 21 | 139 ± 20 |
G114 | –64 ± 4 | –41 ± 3 | – | – |
V115 | –61 ± 3 | –42 ± 5 | – | – |
A116 | –62 ± 5 | –37 ± 4 | –62 ± 4 | –39 ± 2 |
T117 | –68 ± 8 | –38 ± 7 | –68 ± 8 | –39 ± 6 |
A118 | –64 ± 7 | –39 ± 6 | –61 ± 3 | –43 ± 6 |
A119 | –66 ± 4 | –43 ± 4 | –66 ± 8 | –40 ± 7 |
Q120 | –67 ± 6 | –40 ± 7 | –69 ± 6 | –39 ± 8 |
V121 | –68 ± 7 | –41 ± 7 | –64 ± 7 | –41 ± 5 |
T122 | –63 ± 4 | –40 ± 10 | –63 ± 4 | –42 ± 7 |
A123 | –60 ± 5 | –38 ± 7 | –64 ± 4 | –37 ± 6 |
A124 | –64 ± 5 | –39 ± 9 | –64 ± 6 | –45 ± 4 |
V125 | –61 ± 5 | –44 ± 7 | –68 ± 19 | –40 ± 12 |
A126 | –59 ± 4 | –40 ± 6 | – | – |
L127 | –75 ± 19 | –36 ± 14 | – | – |
The DOPC/DOPG values were predicted from the main set of chemical shifts.
Depth of Insertion of FPK4 in the POPC/POPG Membrane
Figure 4
Figure 4. Depth of insertion of FPK4 in the POPC/POPG membrane from gel-phase spin diffusion. (a) Representative 2D spectra with 0 and 25 ms spin diffusion mixing at 258 K. (b) 1H cross sections for the peptide Cα peaks (red) and the lipid CH2 peak (black). Already at 4 ms, the peptide and lipid 1H cross sections have similar intensity patterns, indicating that the peptide is well inserted into the membrane. (c) 13C cross sections extracted from the water (blue) and lipid CH2 (black) 1H chemical shifts from the 4 ms 2D spectra. The N- and C-terminal residues have higher water/lipid intensity ratios than the middle residues. (d) Water/lipid intensity ratios for all labeled sites.
FPK4 Has a Mixed Strand/Helix Conformation in the DOPC/DOPG Membrane
Figure 5
Figure 5. Representative 1D 13C CP MAS spectra of DOPC/DOPG-bound FPK4 as a function of temperature. The VLAAT-FPK4 spectra are shown. At high temperature, mainly β-strand chemical shifts (blue dotted lines) are observed, while at low temperature, both α-helical (red dashed lines) and β-strand chemical shifts are detected.
Figure 6
Figure 6. 2D 13C–13C correlation spectra of DOPC/DOPG-bound FPK4 in the gel phase (233 or 243 K, left column) and the LC phase (303 K, right column). (a, b) GVTAA-FPK4 spectra. (c, d) VLAAT-FPK4 spectra. (e, f) IGALV-FPK4 spectra. (g, h) AAQV-FPK4 spectra. At both temperatures, many residues show mixed α-helical and β-strand chemical shifts.
Depth of Insertion and Lipid Interaction of FPK4 in the DOPC/DOPG Membrane
Figure 7
Figure 7. Depth of insertion of FPK4 in the DOPC/DOPG membrane from gel-phase spin diffusion spectra measured at 243 K. (a) 1H cross sections of the peptide Cα peaks (red) and lipid CH2 peak (black). By 4 ms, the peptide and lipid signals have equilibrated, indicating that the peptide is well inserted into the membrane. (b) 13C cross sections from the water (blue) and lipid CH2 (black) 1H chemical shifts of the 4 ms 2D spectra. The C-terminal α-helical residues have higher water cross peaks than the N-terminal β-strand residues, and the α-helical A123/A124 have higher water cross peaks than the β-strand A123/A124. (c) Water/lipid intensity ratios of all labeled residues in the DOPC/DOPG membrane (blue and red symbols). The β-strand residues have lower water exposure than the α-helical residues. Open symbols indicate the minor conformation. For comparison, the POPC/POPG-bound FPK4 data are also shown (black open symbols).
FPK4 Conformation and Lipid Interaction in the DOPE Membrane
Figure 8
Figure 8. FPK4 interaction with the DOPE membrane. (a, b) Static 31P spectra of the membrane without (a) and with (b) FPK4 from 273 to 313 K. FPK4 increased the Lα-to-HII phase transition temperature and caused a small isotropic peak. (c) 2D 31P–1H correlation spectrum of FPK4-bound DOPE membrane with a spin diffusion mixing time of 225 ms. (d) 1H cross sections from the 2D 31P–1H spectra of peptide-free and peptide-bound DOPE membranes, compared with the 1D 1H single-pulse spectrum (top). The FPK4-bound DOPE membrane has a much weaker water–31P cross peak than the peptide-free membrane.
Figure 9
Figure 9. Conformation and depth of FPK4 in the DOPE membrane. (a) 2D 13C–13C correlation spectrum of a fresh GVTAA-FPK4 sample at 243 K. The peptide exhibits both helix and strand signals. (b) 13C CP-MAS spectra of the initial and equilibrated GVTAA-FPK4 at 246 K. At equilibrium, most residues exhibit β-strand chemical shifts. (c) 100 ms 2D 13C–1H correlation spectrum at 293 K, in the HII phase membrane. Lipid–peptide cross peaks are observed, indicating that the β-strand peptide is inserted into the hydrophobic region of the DOPE membrane.
Figure 10
Figure 10. PIV5 fusion peptide conformations in lipid membranes from solid-state NMR and outside the membrane from crystal structures. (a) Fusion peptide is fully α-helical in POPC/POPG bilayers but adopts a mixed strand/helix conformation in DOPC/DOPG bilayers. The peptide is inserted into both membranes, but the depicted tilt angle is hypothetical. The structures were built using (φ, ψ) torsion angles predicted by TALOS+. (b) Prefusion crystal structures of the PIV5 F protein in the uncleaved (green) (22) and cleaved (red) (21) states. The fusion peptide domain has similar conformations before and after cleavage and has a bend near T117. (c) Prefusion crystal structures of the influenza HA in the uncleaved (green) (76) and cleaved (red) (77) states. The N-terminal half of the fusion peptide is rotated around N12 before and after cleavage. (d) Postfusion crystal structure of the PIV5 F HRA/HRB complex. (14) Seven residues (T122–V128) of the fusion peptide are detected and show α-helical structure extended from HRA. (e) Schematic of the PIV5 fusion peptide conformation in the DOPE membrane. The lipid cylinders and water radius are drawn to scale using 15–18 water molecules per lipid based on the DOPE phase diagram. (65, 66) (f) The hemifusion stalk intermediate showing both negative and positive membrane curvatures and dehydration between two opposing bilayers. Dashed lines indicate the middle of two lipid leaflets.
Discussion
Conformational Polymorphism of the PIV5 Fusion Peptide
Curvature Generation and Membrane Dehydration by the PIV5 Fusion Peptide
Conclusion
Supporting Information
Additional 2D spectra and a table of residue-specific helicity of FPK4. This material is available free of charge via the Internet at http://pubs.acs.org.
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.
Acknowledgment
This work is funded by NIH grant GM066976 to M.H.
References
This article references 86 other publications.
- 1Fields, C. G.; Lloyd, D. H.; Macdonald, R. L.; Ottenson, K. M.; Nobel, R. L. Peptide Res. 1991, 4, 95Google Scholar1HBTU activation for automated Fmoc solid-phase peptide synthesisFields, C. G.; Lloyd, D. H.; Macdonald, R. L.; Otteson, K. M.; Noble, Richard L.Peptide Research (1991), 4 (2), 95-101CODEN: PEREEO; ISSN:1040-5704.Excellent results have been obtained for the 9-fluorenylmethoxycarbonyl (Fmoc) solid-phase syntheses of peptides using the activating reagent 2-(1H-benzotriazol-1-yl)-1,1,3,3-tetramethyluronium hexafluorophosphate (HBTU). Activation occurs very rapidly in DMF and N-methylpyrrolidone, optimal solvents for peptide resin solvation. Complete coupling reactions occur in only 10-30 min. Residues such as arginine, isoleucine, leucine, and valine, which often require double coupling by other activation methods, react with high efficiency by single coupling when HBTU is used. The Fmoc/HBTU chem. has recently been applied to the peptide synthesizers. The incorporation of trityl side-chain protection for Fmoc-Asn and Fmoc-Gln further enhances coupling efficiencies in difficult sequences.
- 2Chang, A.; Dutch, R. E. Viruses 2012, 4, 613Google Scholar2Paramyxovirus fusion and entry: multiple paths to a common endChang, Andres; Dutch, Rebecca E.Viruses (2012), 4 (), 613-636CODEN: VIRUBR; ISSN:1999-4915. (MDPI AG)A review. The paramyxovirus family contains many common human pathogenic viruses, including measles, mumps, the parainfluenza viruses, respiratory syncytial virus, human metapneumovirus, and the zoonotic henipaviruses, Hendra and Nipah. While the expression of a type 1 fusion protein and a type 2 attachment protein is common to all paramyxoviruses, there is considerable variation in viral attachment, the activation and triggering of the fusion protein, and the process of viral entry. In this review, we discuss recent advances in the understanding of paramyxovirus F protein-mediated membrane fusion, an essential process in viral infectivity. We also review the role of the other surface glycoproteins in receptor binding and viral entry, and the implications for viral infection. Throughout, we conc. on the commonalities and differences in fusion triggering and viral entry among the members of the family. Finally, we highlight key unanswered questions and how further studies can identify novel targets for the development of therapeutic treatments against these human pathogens.
- 3Dutch, R. E. PLoS Pathog. 2010, 6, e1000881Google ScholarThere is no corresponding record for this reference.
- 4Lamb, R. A.; Jardetzky, T. S. Curr. Opin. Struct. Biol. 2007, 17, 427Google Scholar4Structural basis of viral invasion: lessons from paramyxovirus FLamb, Robert A.; Jardetzky, Theodore S.Current Opinion in Structural Biology (2007), 17 (4), 427-436CODEN: COSBEF; ISSN:0959-440X. (Elsevier B.V.)A review. The structures of glycoproteins that mediate enveloped virus entry into cells have revealed dramatic structural changes that accompany membrane fusion and provided mechanistic insights into this process. The group of class I viral fusion proteins includes the influenza hemagglutinin, paramyxovirus F, HIV env, and other mechanistically related fusogens, but these proteins are unrelated in sequence and exhibit clearly distinct structural features. Recently detd. crystal structures of the paramyxovirus F protein in two conformations, representing pre-fusion and post-fusion states, reveal a novel protein architecture that undergoes large-scale, irreversible refolding during membrane fusion, extending our understanding of this diverse group of membrane fusion machines.
- 5Russell, C. J.; Luque, L. E. Trends Microbiol. 2006, 14, 243Google ScholarThere is no corresponding record for this reference.
- 6Horvath, C. M.; Lamb, R. A. J. Virol. 1992, 66, 2443Google Scholar6Studies on the fusion peptide of a paramyxovirus fusion glycoprotein: roles of conserved residues in cell fusionHorvath, Curt M.; Lamb, Robert A.Journal of Virology (1992), 66 (4), 2443-55CODEN: JOVIAM; ISSN:0022-538X.The role of residues in the conserved hydrophobic N-terminal fusion peptide of the paramyxovirus fusion (F) protein in causing cell-cell fusion was examd. Mutations were introduced into the cDNA encoding the simian virus 5 (SV5) F protein, the altered F proteins were expressed by using an eukaryotic vector, and their ability to mediate syncytium formation was detd. The mutant F proteins contained both single- and multiple-amino-acid substitutions, and they exhibited a variety of intracellular transport properties and fusion phenotypes. The data indicate that many substitutions in the conserved amino acids of the simian virus 5 F fusion peptide can be tolerated without loss of biol. activity. Mutant F proteins which were not transported to the cell surface did not cause cell-cell fusion, but all of the mutants which were transported to the cell surface were fusion competent, exhibiting fusion properties similar to or better than those of the wild-type F protein. Mutant F proteins contg. glycine-to-alanine substitutions had altered intracellular transport characteristics, yet they exhibited a great increase in fusion activity. The potential structural implications of this substitution and the possible importance of these glycine residues in maintaining appropriate levels of fusion activity are discussed.
- 7Russell, C. J.; Jardetzky, T. S.; Lamb, R. A. J. Virol. 2004, 78, 13727Google Scholar7Conserved glycine residues in the fusion peptide of the paramyxovirus fusion protein regulate activation of the native stateRussell, Charles J.; Jardetzky, Theodore S.; Lamb, Robert A.Journal of Virology (2004), 78 (24), 13727-13742CODEN: JOVIAM; ISSN:0022-538X. (American Society for Microbiology)Hydrophobic fusion peptides (FPs) are the most highly conserved regions of class I viral fusion-mediating glycoproteins (vFGPs). FPs often contain conserved glycine residues thought to be crit. for forming structures that destabilize target membranes. Unexpectedly, a mutation of glycine residues in the FP of the fusion (F) protein from the paramyxovirus simian parainfluenza virus 5 (SV5) resulted in mutant F proteins with hyperactive fusion phenotypes (C. M. Horvath and R. A. Lamb, J. Virol.66:2443-2455, 1992). Here, we constructed G3A and G7A mutations into the F proteins of SV5 (W3A and WR isolates), Newcastle disease virus (NDV), and human parainfluenza virus type 3 (HPIV3). All of the mutant F proteins, except NDV G7A, caused increased cell-cell fusion despite having slight to moderate redns. in cell surface expression compared to those of wild-type F proteins. The G3A and G7A mutations cause SV5 WR F, but not NDV F or HPIV3 F, to be triggered to cause fusion in the absence of coexpression of its homo-typic receptor-binding protein hemagglutinin-neuraminidase (HN), suggesting that NDV and HPIV3 F have stricter requirements for homotypic HN for fusion activation. Dye transfer assays show that the G3A and G7A mutations decrease the energy required to activate F at a step in the fusion cascade preceding prehairpin intermediate formation and hemifusion. Conserved glycine residues in the FP of paramyxovirus F appear to have a primary role in regulating the activation of the metastable native form of F. Glycine residues in the FPs of other class I vFGPs may also regulate fusion activation.
- 8Bissonnette, M. L.; Donald, J. E.; DeGrado, W. F.; Jardetzky, T. S.; Lamb, R. A. J. Mol. Biol. 2009, 386, 14Google ScholarThere is no corresponding record for this reference.
- 9Baquero, E.; Albertini, A. A.; Vachette, P.; Lepault, J.; Bressanelli, S.; Gaudin, Y. Curr. Opin. Virol. 2013, 3, 143Google Scholar9Intermediate conformations during viral fusion glycoprotein structural transitionBaquero, Eduard; Albertini, Aurelie A.; Vachette, Patrice; Lepault, Jean; Bressanelli, Stephane; Gaudin, YvesCurrent Opinion in Virology (2013), 3 (2), 143-150CODEN: COVUAF; ISSN:1879-6257. (Elsevier B. V.)A review. Entry of enveloped viruses into cells requires the fusion of viral and cellular membranes, driven by conformational changes in viral glycoproteins. Three different classes of viral fusion proteins have been hitherto identified based on common structural elements. Crystal structures have provided static pictures of pre-fusion and post-fusion conformations of these proteins and have revealed the dramatic reorganization of the mols., but the transition pathway remains elusive. In this review, we will focus on recent data aiming to characterize intermediate structures during the conformational change. All these data support the existence of a pre-hairpin intermediate, but its oligomeric status is still a matter of debate.
- 10Weissenhorn, W.; Hinz, A.; Gaudin, Y. FEBS Lett. 2007, 581, 2150Google ScholarThere is no corresponding record for this reference.
- 11White, J. M.; Delos, S. E.; Brecher, M.; Schornberg, K. Crit. Rev. Biochem. Mol. Biol. 2008, 43, 189Google Scholar11Structures and Mechanisms of Viral Membrane Fusion Proteins: Multiple Variations on a Common ThemeWhite, Judith M.; Delos, Sue E.; Brecher, Matthew; Schornberg, KathrynCritical Reviews in Biochemistry and Molecular Biology (2008), 43 (3), 189-219CODEN: CRBBEJ; ISSN:1040-9238. (Informa Healthcare)A review. Recent work has identified three distinct classes of viral membrane fusion proteins based on structural criteria. In addn., there are at least four distinct mechanisms by which viral fusion proteins can be triggered to undergo fusion-inducing conformational changes. Viral fusion proteins also contain different types of fusion peptides and vary in their reliance on accessory proteins. These differing features combine to yield a rich diversity of fusion proteins. Yet despite this staggering diversity, all characterized viral fusion proteins convert from a fusion-competent state (dimers or trimers, depending on the class) to a membrane-embedded homotrimeric prehairpin, and then to a trimer-of-hairpins that brings the fusion peptide, attached to the target membrane, and the transmembrane domain, attached to the viral membrane, into close proximity thereby facilitating the union of viral and target membranes. During these conformational conversions, the fusion proteins induce membranes to progress through stages of close apposition, hemifusion, and then the formation of small, and finally large, fusion pores. Clearly, highly divergent proteins have converged on the same overall strategy to mediate fusion, an essential step in the life cycle of every enveloped virus.
- 12Lamb, R. A.; Paterson, R. G.; Jardetzky, T. S. Virology 2006, 344, 30Google ScholarThere is no corresponding record for this reference.
- 13Harrison, S. C. Nat. Struct. Mol. Biol. 2008, 15, 690Google Scholar13Viral membrane fusionHarrison, Stephen C.Nature Structural & Molecular Biology (2008), 15 (7), 690-698CODEN: NSMBCU; ISSN:1545-9993. (Nature Publishing Group)A review. Infection by viruses having lipid-bilayer envelopes proceeds through fusion of the viral membrane with a membrane of the target cell. Viral 'fusion proteins' facilitate this process. They vary greatly in structure, but all seem to have a common mechanism of action, in which a ligand-triggered, large-scale conformational change in the fusion protein is coupled to apposition and merger of the 2 bilayers. The authors describe 3 examples: (1) the influenza virus hemagglutinin; (2) the flavivirus E protein; and (3) the vesicular stomatitis virus G protein, in some detail in order to illustrate the ways in which different structures have evolved to implement this common mechanism. Fusion inhibitors can be effective antiviral agents.
- 14Baker, K. A.; Dutch, R. E.; Lamb, R. A.; Jardetzky, T. S. Mol. Cell 1999, 3, 309Google ScholarThere is no corresponding record for this reference.
- 15Chan, D. C.; Kim, P. S. Cell 1998, 93, 681Google ScholarThere is no corresponding record for this reference.
- 16McLellan, J. S.; Yang, Y.; Graham, B. S.; Kwong, P. D. J. Virol. 2011, 85, 7788Google Scholar16Structure of respiratory syncytial virus fusion glycoprotein in the postfusion conformation reveals preservation of neutralizing epitopesMcLellan, Jason S.; Yang, Yongping; Graham, Barney S.; Kwong, Peter D.Journal of Virology (2011), 85 (15), 7788-7796CODEN: JOVIAM; ISSN:0022-538X. (American Society for Microbiology)Respiratory syncytial virus (RSV) invades host cells via a type I fusion (F) glycoprotein that undergoes dramatic structural rearrangements during the fusion process. Neutralizing monoclonal antibodies, such as 101F, palivizumab, and motavizumab, target two major antigenic sites on the RSV F glycoprotein. The structures of these sites as peptide complexes with motavizumab and 101F have been previously detd., but a structure for the trimeric RSV F glycoprotein ectodomain has remained elusive. To address this issue, we undertook structural and biophys. studies on stable ectodomain constructs. Here, we present the 2.8-Å crystal structure of the trimeric RSV F ectodomain in its postfusion conformation. The structure revealed that the 101F and motavizumab epitopes are present in the postfusion state and that their conformations are similar to those obsd. in the antibody-bound peptide structures. Both antibodies bound the postfusion F glycoprotein with high affinity in surface plasmon resonance expts. Modeling of the antibodies bound to the F glycoprotein predicts that the 101F epitope is larger than the linear peptide and restricted to a single protomer in the trimer, whereas motavizumab likely contacts residues on two protomers, indicating a quaternary epitope. Mechanistically, these results suggest that 101F and motavizumab can bind to multiple conformations of the fusion glycoprotein and can neutralize late in the entry process. The structural preservation of neutralizing epitopes in the postfusion state suggests that this conformation can elicit neutralizing antibodies and serve as a useful vaccine antigen.
- 17Swanson, K.; Wen, X.; Leser, G. P.; Paterson, R. G.; Lamb, R. A.; Jardetzky, T. S. Virology 2010, 402, 372Google ScholarThere is no corresponding record for this reference.
- 18Tan, K.; Liu, J.; Wang, J.; Shen, S.; Lu, M. Proc. Natl. Acad. Sci. U.S.A. 1997, 94, 12303Google ScholarThere is no corresponding record for this reference.
- 19Yin, H. S.; Paterson, R. G.; Wen, X.; Lamb, R. A.; Jardetzky, T. S. Proc. Natl. Acad. Sci. U.S.A. 2005, 102, 9288Google ScholarThere is no corresponding record for this reference.
- 20Zhao, X.; Singh, M.; Malashkevich, V. N.; Kim, P. S. Proc. Natl. Acad. Sci. U.S.A. 2000, 97, 14172Google ScholarThere is no corresponding record for this reference.
- 21Welch, B. D.; Liu, Y.; Kors, C. A.; Leser, G. P.; Jardetzky, T. S.; Lamb, R. A. Proc. Natl. Acad. Sci. U.S.A. 2012, 109, 16672Google ScholarThere is no corresponding record for this reference.
- 22Yin, H. S.; Wen, X.; Paterson, R. G.; Lamb, R. A.; Jardetzky, T. S. Nature 2006, 439, 38Google ScholarThere is no corresponding record for this reference.
- 23Kim, Y. H.; Donald, J. E.; Grigoryan, G.; Leser, G. P.; Fadeev, A. Y.; Lamb, R. A.; DeGrado, W. F. Proc. Natl. Acad. Sci. U.S.A. 2011, 108, 20992Google ScholarThere is no corresponding record for this reference.
- 24Han, X.; Bushweller, J. H.; Cafiso, D. S.; Tamm, L. K. Nat. Struct. Biol. 2001, 8, 715Google ScholarThere is no corresponding record for this reference.
- 25Jaroniec, C. P.; Kaufman, J. D.; Stahl, S. J.; Viard, M.; Blumenthal, R.; Wingfield, P. T.; Bax, A. Biochemistry 2005, 44, 16167Google ScholarThere is no corresponding record for this reference.
- 26Lai, A. L.; Tamm, L. K. J. Biol. Chem. 2010, 285, 37467Google ScholarThere is no corresponding record for this reference.
- 27Li, Y.; Han, X.; Lai, A. L.; Bushweller, J. H.; Cafiso, D. S.; Tamm, L. K. J. Virol. 2005, 79, 12065Google Scholar27Membrane structures of the hemifusion-inducing fusion peptide: mutant G1S and the fusion-blocking mutant G1V of influenza virus hemagglutinin suggest a mechanism for pore opening in membrane fusionLi, Yinling; Han, Xing; Lai, Alex L.; Bushweller, John H.; Cafiso, David S.; Tamm, Lukas K.Journal of Virology (2005), 79 (18), 12065-12076CODEN: JOVIAM; ISSN:0022-538X. (American Society for Microbiology)Influenza virus hemagglutinin (HA)-mediated membrane fusion is initiated by a conformational change that releases a V-shaped hydrophobic fusion domain, the fusion peptide, into the lipid bilayer of the target membrane. The most N-terminal residue of this domain, a glycine, is highly conserved and is particularly crit. for HA function; G1S and G1V mutant HAs cause hemifusion and abolish fusion, resp. We have detd. the at. resoln. structures of the G1S and G1V mutant fusion domains in membrane environments. G1S forms a V with a disrupted "glycine edge" on its N-terminal arm and G1V adopts a slightly tilted linear helical structure in membranes. Abolishment of the kink in G1V results in reduced hydrophobic penetration of the lipid bilayer and an increased propensity to form β-structures at the membrane surface. These results underline the functional importance of the kink in the fusion peptide and suggest a structural role for the N-terminal glycine ridge in viral membrane fusion.
- 28Li, Y.; Tamm, L. K. Biophys. J. 2007, 93, 876Google ScholarThere is no corresponding record for this reference.
- 29Lorieau, J. L.; Louis, J. M.; Bax, A. Proc. Natl. Acad. Sci. U.S.A. 2010, 107, 11341Google ScholarThere is no corresponding record for this reference.
- 30Lorieau, J. L.; Louis, J. M.; Schwieters, C. D.; Bax, A. Proc. Natl. Acad. Sci. U.S.A. 2012, 109, 19994Google ScholarThere is no corresponding record for this reference.
- 31Lai, A. L.; Tamm, L. K. J. Biol. Chem. 2007, 282, 23946Google ScholarThere is no corresponding record for this reference.
- 32Tamm, L. K.; Lai, A. L.; Li, Y. Biochim. Biophys. Acta 2007, 1768, 3052Google ScholarThere is no corresponding record for this reference.
- 33Epand, R. M. Biochim. Biophys. Acta 2003, 1614, 116Google ScholarThere is no corresponding record for this reference.
- 34Li, Y.; Han, X.; Tamm, L. K. Biochemistry 2003, 42, 7245Google ScholarThere is no corresponding record for this reference.
- 35Sun, Y.; Weliky, D. P. J. Am. Chem. Soc. 2009, 131, 13228Google ScholarThere is no corresponding record for this reference.
- 36Lorieau, J. L.; Louis, J. M.; Bax, A. Biopolymers 2013, 99, 189Google ScholarThere is no corresponding record for this reference.
- 37Gordon, L. M.; Mobley, P. W.; Lee, W.; Eskandari, S.; Kaznessis, Y. N.; Sherman, M. A.; Waring, A. J. Protein Sci. 2004, 13, 1012Google Scholar37Conformational mapping of the N-terminal peptide of HIV-1 gp41 in lipid detergent and aqueous environments using 13C-enhanced Fourier transform infrared spectroscopyGordon, Larry M.; Mobley, Patrick W.; Lee, William; Eskandari, Sepehr; Kaznessis, Yiannis N.; Sherman, Mark A.; Waring, Alan J.Protein Science (2004), 13 (4), 1012-1030CODEN: PRCIEI; ISSN:0961-8368. (Cold Spring Harbor Laboratory Press)The N-terminal domain of HIV-1 glycoprotein 41,000 (gp41) participates in viral fusion processes. Here, we use phys. and computational methodologies to examine the secondary structure of a peptide based on the N terminus (FP; residues 1-23) in aq. and detergent environments. 12C-Fourier transform IR (FTIR) spectroscopy indicated greater α-helix for FP in lipid-detergent SDS and aq. phosphate-buffered saline (PBS) than in only PBS. 12C-FTIR spectra also showed disordered FP conformations in these two environments, along with substantial β-structure for FP alone in PBS. In expts. that map conformations to specific residues, isotope-enhanced FTIR spectroscopy was performed using FP peptides labeled with 13C-carbonyl. 13C-FTIR results on FP in SDS at low peptide loading indicated α-helix (residues 5 to 16) and disordered conformations (residues 1-4). Because earlier 13C-FTIR anal. of FP in lipid bilayers demonstrated α-helix for residues 1-16 at low peptide loading, the FP structure in SDS micelles only approximates that found for FP with membranes. Mol. dynamics simulations of FP in an explicit SDS micelle indicate that the fraying of the first three to four residues may be due to the FP helix moving to one end of the micelle. In PBS alone, however, electron microscopy of FP showed large fibrils, while 13C-FTIR spectra demonstrated antiparallel β-sheet for FP (residues 1-12), analogous to that reported for amyloid peptides. Because FP and amyloid peptides each exhibit plaque formation, α-helix to β-sheet interconversion, and membrane fusion activity, amyloid and N-terminal gp41 peptides may belong to the same superfamily of proteins.
- 38Qiang, W.; Bodner, M. L.; Weliky, D. P. J. Am. Chem. Soc. 2008, 130, 5459Google ScholarThere is no corresponding record for this reference.
- 39Qiang, W.; Sun, Y.; Weliky, D. P. Proc. Natl. Acad. Sci. U.S.A. 2009, 106, 15314Google ScholarThere is no corresponding record for this reference.
- 40Lai, A. L.; Moorthy, A. E.; Li, Y.; Tamm, L. K. J. Mol. Biol. 2012, 418, 3Google ScholarThere is no corresponding record for this reference.
- 41Gordon, L. M.; Mobley, P. W.; Pilpa, R.; Sherman, M. A.; Waring, A. J. Biochim. Biophys. Acta 2002, 1559, 96Google ScholarThere is no corresponding record for this reference.
- 42Rafalski, M.; Lear, J. D.; DeGrado, W. F. Biochemistry 1990, 29, 7917Google ScholarThere is no corresponding record for this reference.
- 43Saez-Cirion, A.; Nieva, J. L. Biochim. Biophys. Acta 2002, 1564, 57Google ScholarThere is no corresponding record for this reference.
- 44Gabrys, C. M.; Qiang, W.; Sun, Y.; Xie, L.; Schmick, S. D.; Weliky, D. P. J. Phys. Chem. A 2013, 117, 9848Google ScholarThere is no corresponding record for this reference.
- 45Hong, M.; Zhang, Y.; Hu, F. Annu. Rev. Phys. Chem. 2012, 63, 1Google ScholarThere is no corresponding record for this reference.
- 46Fuhrmans, M.; Marrink, S. J. J. Am. Chem. Soc. 2012, 134, 1543Google ScholarThere is no corresponding record for this reference.
- 47Siegel, D. P. Biophys. J. 1999, 76, 291Google ScholarThere is no corresponding record for this reference.
- 48Kasson, P. M.; Pande, V. S. PLoS Comput. Biol. 2007, 3, e220Google ScholarThere is no corresponding record for this reference.
- 49Chernomordik, L. V.; Kozlov, M. M. Cell 2005, 123, 375Google ScholarThere is no corresponding record for this reference.
- 50Tamm, L. K.; Han, X. Biosci. Rep. 2000, 20, 501Google Scholar50Viral fusion peptides: a tool set to disrupt and connect biological membranesTamm, Lukas K.; Han, XingBioscience Reports (2000), 20 (6), 501-518CODEN: BRPTDT; ISSN:0144-8463. (Kluwer Academic/Plenum Publishers)A review, with 56 refs., on the structure and function of viral fusion peptides. The fusion peptides of influenza virus hemagglutinin and human immunodeficiency virus are used as paradigms. Fusion peptides assocd. with lipid bilayers are conformationally polymorphic. Current evidence suggests that the fusion-promoting state is the obliquely inserted α-helix. Fusion peptides also have a tendency to self-assoc. into β-sheets at membrane surfaces. Although the conformational conversion between α- and β-states is reversible under controlled conditions, its physiol. relevance is not yet known. The energetics of peptide insertion and self-assocn. could be measured recently using more sol. 2nd generation fusion peptides. Fusion peptides have been reported to change membrane curvature and the state of hydration of membrane surfaces. The combined results are built into a model for the mechanism by which fusion peptides are proposed to assist in biol. membrane fusion.
- 51Siegel, D. P.; Epand, R. M. Biochim. Biophys. Acta 2000, 1468, 87Google ScholarThere is no corresponding record for this reference.
- 52Yao, H.; Hong, M. J. Mol. Biol. 2013, 425, 563Google ScholarThere is no corresponding record for this reference.
- 53Takegoshi, K.; Nakamura, S.; Terao, T. Chem. Phys. Lett. 2001, 344, 631Google ScholarThere is no corresponding record for this reference.
- 54Hong, M.; Griffin, R. G. J. Am. Chem. Soc. 1998, 120, 7113Google ScholarThere is no corresponding record for this reference.
- 55Huster, D.; Yao, X. L.; Hong, M. J. Am. Chem. Soc. 2002, 124, 874Google ScholarThere is no corresponding record for this reference.
- 56Mani, R.; Cady, S. D.; Tang, M.; Waring, A. J.; Lehrer, R. I.; Hong, M. Proc. Natl. Acad. Sci. U.S.A. 2006, 103, 16242Google ScholarThere is no corresponding record for this reference.
- 57Wang, T.; Yao, H.; Hong, M. J. Biomol. NMR 2013, 56, 139Google ScholarThere is no corresponding record for this reference.
- 58Donald, J. E.; Zhang, Y.; Fiorin, G.; Carnevale, V.; Slochower, D. R.; Gai, F.; Klein, M. L.; DeGrado, W. F. Proc. Natl. Acad. Sci. U.S.A. 2011, 108, 3958Google ScholarThere is no corresponding record for this reference.
- 59Shen, Y.; Delaglio, F.; Cornilescu, G.; Bax, A. J. Biomol. NMR 2009, 44, 213Google ScholarThere is no corresponding record for this reference.
- 60Huster, D.; Yao, X.; Hong, M. J. Am. Chem. Soc. 2002, 124, 874Google ScholarThere is no corresponding record for this reference.
- 61Cady, S. D.; Goodman, C.; Tatko, C.; DeGrado, W. F.; Hong, M. J. Am. Chem. Soc. 2007, 129, 5719Google ScholarThere is no corresponding record for this reference.
- 62Hong, M.; Doherty, T. Chem. Phys. Lett. 2006, 432, 296Google ScholarThere is no corresponding record for this reference.
- 63Drechsler, A.; Anderluh, G.; Norton, R. S.; Separovic, F. Biochim. Biophys. Acta 2010, 1798, 244Google ScholarThere is no corresponding record for this reference.
- 64Traikia, M.; Warschawski, D. E.; Recouvreur, M.; Cartaud, J.; Devaux, P. F. Eur. Biophys. J. 2000, 29, 184Google ScholarThere is no corresponding record for this reference.
- 65Gawrisch, K.; Parsegian, V. A.; Hajduk, D. A.; Tate, M. W.; Graner, S. M.; Fuller, N. L.; Rand, R. P. Biochemistry 1992, 31, 2856Google ScholarThere is no corresponding record for this reference.
- 66Rand, R. P.; Fuller, N. L. Biophys. J. 1994, 66, 2127Google ScholarThere is no corresponding record for this reference.
- 67Thayer, A. M.; Kohler, S. J. Biochemistry 1981, 20, 6831Google ScholarThere is no corresponding record for this reference.
- 68Tenchov, B. G.; MacDonald, R. C.; Lentz, B. R. Biophys. J. 2013, 104, 1029Google ScholarThere is no corresponding record for this reference.
- 69Schmidt, N.; Mishra, A.; Lai, G. H.; Wong, G. C. FEBS Lett. 2010, 584, 1806Google ScholarThere is no corresponding record for this reference.
- 70Mishra, A.; Gordon, V. D.; Yang, L.; Coridan, R.; Wong, G. C. L. Angew. Chem., Int. Ed. 2008, 47, 2986Google ScholarThere is no corresponding record for this reference.
- 71Kučerka, N.; Nieh, M. P.; Katsaras, J. Biochim. Biophys. Acta 2011, 1808, 2761Google ScholarThere is no corresponding record for this reference.
- 72Kucerka, N.; Tristram-Nagle, S.; Nagle, J. F. J. Membr. Biol. 2005, 208, 193Google Scholar72Structure of Fully Hydrated Fluid Phase Lipid Bilayers with Monounsaturated ChainsKucerka, Norbert; Tristram-Nagle, Stephanie; Nagle, John F.Journal of Membrane Biology (2006), 208 (3), 193-202CODEN: JMBBBO; ISSN:0022-2631. (Springer)Quant. structures are obtained at 30°C for the fully hydrated fluid phases of palmitoyloleoylphosphatidylcholine (POPC), with a double bond on the sn-2 hydrocarbon chain, and for dierucoylphosphatidylcholine (di22:1PC), with a double bond on each hydrocarbon chain. The form factors F(qz) for both lipids are obtained using a combination of 3 methods: (1) Volumetric measurements provide F(0), (2) x-ray scattering from extruded unilamellar vesicles provides |F(qz)| for low qz, (3) Diffuse x-ray scattering from oriented stacks of bilayers provides |F(qz)| for high qz. Also, data using method (2) are added to the authors' recent data for dioleoylphosphatidylcholine (DOPC) using methods (1) and (3); the new DOPC data agree very well with the recent data and with (4) the authors' older data obtained using a liq. crystallog. x-ray method. The authors used hybrid electron d. models to obtain structural results from these form factors. The result for area per lipid (A) for DOPC 72.4 ± 0.5 Å2 agrees well with the authors' earlier publications, and the authors find A = 69.3 ± 0.5 Å2 for di22:1PC and A = 68.3 ± 1.5 Å2 for POPC. The authors obtain the values for 5 different av. thicknesses: hydrophobic, steric, head-head, phosphate-phosphate and Luzzati. Comparison of the results for these 3 lipids and for the authors' recent dimyristoylphosphatidylcholine (DMPC) detn. provides quant. measures of the effect of unsatn. on bilayer structure. The authors' results suggest that lipids with one monounsatd. chain have quant. bilayer structures closer to lipids with 2 monounsatd. chains than to lipids with 2 completely satd. chains.
- 73Chernomordik, L. V.; Kozlov, M. M. Annu. Rev. Biochem. 2003, 72, 175Google Scholar73Protein-lipid interplay in fusion and fission of biological membranesChernomordik, Leonid V.; Kozlov, Michael M.Annual Review of Biochemistry (2003), 72 (), 175-207CODEN: ARBOAW; ISSN:0066-4154. (Annual Reviews Inc.)A review. Disparate biol. processes involve fusion of two membranes into one and fission of one membrane into two. To formulate the possible job description for the proteins that mediate remodeling of biol. membranes, we analyze the energy price of disruption and bending of membrane lipid bilayers at the different stages of bilayer fusion. The phenomenol. and the pathways of the well-characterized reactions of biol. remodeling, such as fusion mediated by influenza hemagglutinin, are compared with those studied for protein-free bilayers. We briefly consider some proteins involved in fusion and fission, and the dependence of remodeling on the lipid compn. of the membranes. The specific hypothetical mechanisms by which the proteins can lower the energy price of the bilayer rearrangement are discussed in light of the exptl. data and the requirements imposed by the elastic properties of the bilayer.
- 74Markvoort, A. J.; Marrink, S. J. Curr. Top. Membr. 2011, 68, 259Google Scholar74Lipid acrobatics in the membrane fusion arenaMarkvoort, Albert J.; Marrink, Siewert J.Current Topics in Membranes (2011), 68 (Membrane Fusion), 259-294CODEN: CTMEET; ISSN:1063-5823. (Elsevier Inc.)A review on the recent contribution of computer simulation approaches to unravel the mol. details of membrane fusion. It considers particle-based simulations of apposed bilayers, bilayers and vesicles, or between vesicles. It sets out with a short overview of the historical background of fusion-related simulation studies, introducing the main particle-based simulation techniques used. It proceeds with a detailed description of the various fusion pathways that are obsd. in these simulations, and the energetics involved. Special attention is devoted to discuss the accumulating evidence that the first barrier to fusion is in fact the splaying of a single lipid tail. Subsequently, the process of vesicle fission is described, showing that it is not just fusion reversed. Finally, the growing body of simulations probing the effect of fusion peptides and proteins on the fusion process is discussed.
- 75Chernomordik, L. V.; Leikina, E.; Frolov, V.; Bronk, P.; Zimmerberg, J. J. Cell Biol. 1997, 136, 81Google ScholarThere is no corresponding record for this reference.
- 76Chen, J.; Lee, K. H.; Steinhauer, D. A.; Stevens, D. J.; Skehel, J. J.; Wiley, D. C. Cell 1998, 95, 409Google ScholarThere is no corresponding record for this reference.
- 77Weis, W. I.; Brunger, A. T.; Skehel, J. J.; Wiley, D. C. J. Mol. Biol. 1990, 212, 737Google ScholarThere is no corresponding record for this reference.
- 78Smith, E. C.; Gregory, S. M.; Tamm, L. K.; Creamer, T. P.; Dutch, R. E. J. Biol. Chem. 2012, 287, 30035Google ScholarThere is no corresponding record for this reference.
- 79Epand, R. M.; Epand, R. F. Biochem. Biophys. Res. Commun. 1994, 202, 1420Google ScholarThere is no corresponding record for this reference.
- 80Epand, R. M.; Epand, R. F.; Martin, I.; Ruysschaert, J. M. Biochemistry 2001, 40, 8800Google ScholarThere is no corresponding record for this reference.
- 81Gabrys, C. M.; Yang, R.; Wasniewski, C. M.; Yang, J.; Canlas, C. G.; Qiang, W.; Sun, Y.; Weliky, D. P. Biochim. Biophys. Acta 2010, 1798, 194Google ScholarThere is no corresponding record for this reference.
- 82Pereira, F. B.; Valpuesta, J. M.; Basañez, G.; Goñi, F. M.; Nieva, J. L. Chem. Phys. Lipids 1999, 103, 11Google ScholarThere is no corresponding record for this reference.
- 83Tristram-Nagle, S.; Chan, R.; Kooijman, E.; Uppamoochikkal, P.; Qiang, W.; Weliky, D. P.; Nagle, J. F. J. Mol. Biol. 2010, 402, 139Google ScholarThere is no corresponding record for this reference.
- 84Kaasgaard, T.; Drummond, C. J. Phys. Chem. Chem. Phys. 2006, 8, 4957Google ScholarThere is no corresponding record for this reference.
- 85Kulkarni, C. V.; Wachter, W.; Iglesias-Salto, G.; Engelskirchen, S.; Ahualli, S. Phys. Chem. Chem. Phys. 2011, 13, 3004Google ScholarThere is no corresponding record for this reference.
- 86Zhang, H.; Neal, S.; Wishart, D. S. J. Biomol. NMR 2003, 25, 173Google ScholarThere is no corresponding record for this reference.
Cited By
Smart citations by scite.ai include citation statements extracted from the full text of the citing article. The number of the statements may be higher than the number of citations provided by ACS Publications if one paper cites another multiple times or lower if scite has not yet processed some of the citing articles.
This article is cited by 44 publications.
- Mariana Valério, Diogo A. Mendonça, João Morais, Carolina C. Buga, Carlos H. Cruz, Miguel A.R.B. Castanho, Manuel N. Melo, Cláudio M. Soares, Ana Salomé Veiga, Diana Lousa. Parainfluenza Fusion Peptide Promotes Membrane Fusion by Assembling into Oligomeric Porelike Structures. ACS Chemical Biology 2022, 17
(7)
, 1831-1843. https://doi.org/10.1021/acschembio.2c00208
- Ujjayini Ghosh, David P. Weliky. Rapid 2H NMR Transverse Relaxation of Perdeuterated Lipid Acyl Chains of Membrane with Bound Viral Fusion Peptide Supports Large-Amplitude Motions of These Chains That Can Catalyze Membrane Fusion. Biochemistry 2021, 60
(35)
, 2637-2651. https://doi.org/10.1021/acs.biochem.1c00316
- Wenting Wang, Junjun Tan, Shuji Ye. Unsaturated Lipid Accelerates Formation of Oligomeric β-Sheet Structure of GP41 Fusion Peptide in Model Cell Membrane. The Journal of Physical Chemistry B 2020, 124
(25)
, 5169-5176. https://doi.org/10.1021/acs.jpcb.0c02464
- Sabine Reißer, Erik Strandberg, Thomas Steinbrecher, Marcus Elstner, Anne S. Ulrich. Best of Two Worlds? How MD Simulations of Amphiphilic Helical Peptides in Membranes Can Complement Data from Oriented Solid-State NMR. Journal of Chemical Theory and Computation 2018, 14
(11)
, 6002-6014. https://doi.org/10.1021/acs.jctc.8b00283
- Byungsu Kwon, Myungwoon Lee, Alan J. Waring, Mei Hong. Oligomeric Structure and Three-Dimensional Fold of the HIV gp41 Membrane-Proximal External Region and Transmembrane Domain in Phospholipid Bilayers. Journal of the American Chemical Society 2018, 140
(26)
, 8246-8259. https://doi.org/10.1021/jacs.8b04010
- Hongwei Yao, Myungwoon Lee, Shu-Yu Liao, and Mei Hong . Solid-State Nuclear Magnetic Resonance Investigation of the Structural Topology and Lipid Interactions of a Viral Fusion Protein Chimera Containing the Fusion Peptide and Transmembrane Domain. Biochemistry 2016, 55
(49)
, 6787-6800. https://doi.org/10.1021/acs.biochem.6b00568
- Tuo Wang and Mei Hong . Investigation of the Curvature Induction and Membrane Localization of the Influenza Virus M2 Protein Using Static and Off-Magic-Angle Spinning Solid-State Nuclear Magnetic Resonance of Oriented Bicelles. Biochemistry 2015, 54
(13)
, 2214-2226. https://doi.org/10.1021/acs.biochem.5b00127
- Mariana Valério, Carolina C. Buga, Manuel N. Melo, Cláudio M. Soares, Diana Lousa. Viral entry mechanisms: the role of molecular simulation in unlocking a key step in viral infections. FEBS Open Bio 2025, 15
(2)
, 269-284. https://doi.org/10.1002/2211-5463.13908
- Lifen Zheng, Shenlin Wang. Recent advances in solid‐state nuclear magnetic resonance studies on membrane fusion proteins. The FEBS Journal 2025, 292
(3)
, 483-499. https://doi.org/10.1111/febs.17313
- Jiayu Li, David A. Eagles, Isaac J. Tucker, Anneka C. Pereira Schmidt, Evelyne Deplazes. Secondary structure propensities of the Ebola delta peptide E40 in solution and model membrane environments. Biophysical Chemistry 2024, 314 , 107318. https://doi.org/10.1016/j.bpc.2024.107318
- Xiaoxuan Zheng, Zijian Ni, Quanbing Pei, Mengmeng Wang, Junjun Tan, Shiyu Bai, Fangwen Shi, Shuji Ye. Probing the Molecular Structure and Dynamics of Membrane‐Bound Proteins during Misfolding Processes by Sum‐Frequency Generation Vibrational Spectroscopy. ChemPlusChem 2024, 89
(6)
https://doi.org/10.1002/cplu.202300684
- Nadia El Mammeri, Olivia Gampp, Pu Duan, Mei Hong. Membrane-induced tau amyloid fibrils. Communications Biology 2023, 6
(1)
https://doi.org/10.1038/s42003-023-04847-6
- Farshad C. Azimi, Trevor T. Dean, Karine Minari, Luis G. M. Basso, Tyler D. R. Vance, Vitor Hugo B. Serrão. A Frame-by-Frame Glance at Membrane Fusion Mechanisms: From Viral Infections to Fertilization. Biomolecules 2023, 13
(7)
, 1130. https://doi.org/10.3390/biom13071130
- Aurelio J. Dregni, Matthew J. McKay, Wahyu Surya, Maria Queralt-Martin, João Medeiros-Silva, Harrison K. Wang, Vicente Aguilella, Jaume Torres, Mei Hong. The Cytoplasmic Domain of the SARS-CoV-2 Envelope Protein Assembles into a β-Sheet Bundle in Lipid Bilayers. Journal of Molecular Biology 2023, 435
(5)
, 167966. https://doi.org/10.1016/j.jmb.2023.167966
- Kayano Izumi, Batsaikhan Mijiddorj, Chihiro Saito, Izuru Kawamura, Ryuji Kawano. Peptide Structure: Electrophysiological Analysis, Nuclear Magnetic Resonance Analysis, and Molecular Dynamics Simulation of Direct Penetration of Cell‐Penetrating Peptides through Bilayer Lipid Membranes. 2023, 109-140. https://doi.org/10.1002/9783527835997.ch8
- Florian Turbant, Jehan Waeytens, Camille Campidelli, Marianne Bombled, Denis Martinez, Axelle Grélard, Birgit Habenstein, Vincent Raussens, Marisela Velez, Frank Wien, Véronique Arluison. Unraveling Membrane Perturbations Caused by the Bacterial Riboregulator Hfq. International Journal of Molecular Sciences 2022, 23
(15)
, 8739. https://doi.org/10.3390/ijms23158739
- Ankita Joardar, Gourab Prasad Pattnaik, Hirak Chakraborty. Mechanism of Membrane Fusion: Interplay of Lipid and Peptide. The Journal of Membrane Biology 2022, 255
(2-3)
, 211-224. https://doi.org/10.1007/s00232-022-00233-1
- Luis Guilherme Mansor Basso, Ana Eliza Zeraik, Ana Paula Felizatti, Antonio José Costa-Filho. Membranotropic and biological activities of the membrane fusion peptides from SARS-CoV spike glycoprotein: The importance of the complete internal fusion peptide domain. Biochimica et Biophysica Acta (BBA) - Biomembranes 2021, 1863
(11)
, 183697. https://doi.org/10.1016/j.bbamem.2021.183697
- Madeleine Sutherland, Byungsu Kwon, Mei Hong. Interactions of HIV gp41's membrane-proximal external region and transmembrane domain with phospholipid membranes from 31P NMR. Biochimica et Biophysica Acta (BBA) - Biomembranes 2021, 1863
(11)
, 183723. https://doi.org/10.1016/j.bbamem.2021.183723
- Alex L. Lai, Jack H. Freed. SARS-CoV-2 Fusion Peptide has a Greater Membrane Perturbating Effect than SARS-CoV with Highly Specific Dependence on Ca2+. Journal of Molecular Biology 2021, 433
(10)
, 166946. https://doi.org/10.1016/j.jmb.2021.166946
- Tara C. Marcink, Matteo Porotto, Anne Moscona. Parainfluenza virus entry at the onset of infection. 2021, 1-29. https://doi.org/10.1016/bs.aivir.2021.07.001
- Ujjayini Ghosh, David P. Weliky. 2H nuclear magnetic resonance spectroscopy supports larger amplitude fast motion and interference with lipid chain ordering for membrane that contains β sheet human immunodeficiency virus gp41 fusion peptide or helical hairpin influenza virus hemagglutinin fusion peptide at fusogenic pH. Biochimica et Biophysica Acta (BBA) - Biomembranes 2020, 1862
(10)
, 183404. https://doi.org/10.1016/j.bbamem.2020.183404
- Hirak Chakraborty, Surajit Bhattacharjya. Mechanistic insights of host cell fusion of SARS-CoV-1 and SARS-CoV-2 from atomic resolution structure and membrane dynamics. Biophysical Chemistry 2020, 265 , 106438. https://doi.org/10.1016/j.bpc.2020.106438
- Byungsu Kwon, Taraknath Mandal, Matthew R. Elkins, Younghoon Oh, Qiang Cui, Mei Hong. Cholesterol Interaction with the Trimeric HIV Fusion Protein gp41 in Lipid Bilayers Investigated by Solid-State NMR Spectroscopy and Molecular Dynamics Simulations. Journal of Molecular Biology 2020, 432
(16)
, 4705-4721. https://doi.org/10.1016/j.jmb.2020.06.017
- Xue Kang, Christopher Elson, Jackson Penfield, Alex Kirui, Adrian Chen, Liqun Zhang, Tuo Wang. Integrated solid-state NMR and molecular dynamics modeling determines membrane insertion of human β-defensin analog. Communications Biology 2019, 2
(1)
https://doi.org/10.1038/s42003-019-0653-6
- Anthony Legrand, Denis Martinez, Axelle Grélard, Melanie Berbon, Estelle Morvan, Arpita Tawani, Antoine Loquet, Sébastien Mongrand, Birgit Habenstein. Nanodomain Clustering of the Plant Protein Remorin by Solid-State NMR. Frontiers in Molecular Biosciences 2019, 6 https://doi.org/10.3389/fmolb.2019.00107
- Myungwoon Lee, Chloe A. Morgan, Mei Hong. Fully hydrophobic HIV gp41 adopts a hemifusion-like conformation in phospholipid bilayers. Journal of Biological Chemistry 2019, 294
(40)
, 14732-14744. https://doi.org/10.1074/jbc.RA119.009542
- Shu Y. Liao, Myungwoon Lee, Mei Hong. Interplay between membrane curvature and protein conformational equilibrium investigated by solid-state NMR. Journal of Structural Biology 2019, 206
(1)
, 20-28. https://doi.org/10.1016/j.jsb.2018.02.007
- Rong Han, Yufei Yang, Shenlin Wang. Longitudinal Relaxation Optimization Enhances
1
H‐Detected HSQC in Solid‐State NMR Spectroscopy on Challenging Biological Systems. Chemistry – A European Journal 2019, 25
(16)
, 4115-4122. https://doi.org/10.1002/chem.201805327
- Catarina M. Morais, Ana M. Cardoso, Pedro P. Cunha, Luísa Aguiar, Nuno Vale, Emílio Lage, Marina Pinheiro, Cláudia Nunes, Paula Gomes, Salette Reis, M. Margarida C.A. Castro, Maria C. Pedroso de Lima, Amália S. Jurado. Acylation of the S413-PV cell-penetrating peptide as a means of enhancing its capacity to mediate nucleic acid delivery: Relevance of peptide/lipid interactions. Biochimica et Biophysica Acta (BBA) - Biomembranes 2018, 1860
(12)
, 2619-2634. https://doi.org/10.1016/j.bbamem.2018.10.002
- Michelle W. Lee, Ernest Y. Lee, Andrew L. Ferguson, Gerard C.L. Wong. Machine learning antimicrobial peptide sequences: Some surprising variations on the theme of amphiphilic assembly. Current Opinion in Colloid & Interface Science 2018, 38 , 204-213. https://doi.org/10.1016/j.cocis.2018.11.003
- Venkata S. Mandala, Jonathan K. Williams, Mei Hong. Structure and Dynamics of Membrane Proteins from Solid-State NMR. Annual Review of Biophysics 2018, 47
(1)
, 201-222. https://doi.org/10.1146/annurev-biophys-070816-033712
- Myungwoon Lee, Hongwei Yao, Byungsu Kwon, Alan J. Waring, Peter Ruchala, Chandan Singh, Mei Hong. Conformation and Trimer Association of the Transmembrane Domain of the Parainfluenza Virus Fusion Protein in Lipid Bilayers from Solid-State NMR: Insights into the Sequence Determinants of Trimer Structure and Fusion Activity. Journal of Molecular Biology 2018, 430
(5)
, 695-709. https://doi.org/10.1016/j.jmb.2018.01.002
- Stefanie Schrottke, Anette Kaiser, Gerrit Vortmeier, Sylvia Els-Heindl, Dennis Worm, Mathias Bosse, Peter Schmidt, Holger A. Scheidt, Annette G. Beck-Sickinger, Daniel Huster. Expression, Functional Characterization, and Solid-State NMR Investigation of the G Protein-Coupled GHS Receptor in Bilayer Membranes. Scientific Reports 2017, 7
(1)
https://doi.org/10.1038/srep46128
- Larisa Kordyukova. Structural and functional specificity of Influenza virus haemagglutinin and paramyxovirus fusion protein anchoring peptides. Virus Research 2017, 227 , 183-199. https://doi.org/10.1016/j.virusres.2016.09.014
- Luis G. M. Basso, Eduardo F. Vicente, Edson Crusca, Eduardo M. Cilli, Antonio J. Costa-Filho. SARS-CoV fusion peptides induce membrane surface ordering and curvature. Scientific Reports 2016, 6
(1)
https://doi.org/10.1038/srep37131
- A. Agopian, M. Quetin, S. Castano. Structure and interaction with lipid membrane models of Semliki Forest virus fusion peptide. Biochimica et Biophysica Acta (BBA) - Biomembranes 2016, 1858
(11)
, 2671-2680. https://doi.org/10.1016/j.bbamem.2016.07.003
- Sean T. Smrt, Justin L. Lorieau. Membrane Fusion and Infection of the Influenza Hemagglutinin. 2016, 37-54. https://doi.org/10.1007/5584_2016_174
- Tzong-Hsien Lee, Daniel J. Hirst, Marie-Isabel Aguilar. New insights into the molecular mechanisms of biomembrane structural changes and interactions by optical biosensor technology. Biochimica et Biophysica Acta (BBA) - Biomembranes 2015, 1848
(9)
, 1868-1885. https://doi.org/10.1016/j.bbamem.2015.05.012
- Hongwei Yao, Michelle W. Lee, Alan J. Waring, Gerard C. L. Wong, Mei Hong. Viral fusion protein transmembrane domain adopts β-strand structure to facilitate membrane topological changes for virus–cell fusion. Proceedings of the National Academy of Sciences 2015, 112
(35)
, 10926-10931. https://doi.org/10.1073/pnas.1501430112
- Lihui Jia, Shuang Liang, Kelly Sackett, Li Xie, Ujjayini Ghosh, David P. Weliky. REDOR solid-state NMR as a probe of the membrane locations of membrane-associated peptides and proteins. Journal of Magnetic Resonance 2015, 253 , 154-165. https://doi.org/10.1016/j.jmr.2014.12.020
- Caitlin M. Quinn, Manman Lu, Christopher L. Suiter, Guangjin Hou, Huilan Zhang, Tatyana Polenova. Magic angle spinning NMR of viruses. Progress in Nuclear Magnetic Resonance Spectroscopy 2015, 86-87 , 21-40. https://doi.org/10.1016/j.pnmrs.2015.02.003
- Sean T. Smrt, Adrian W. Draney, Justin L. Lorieau. The Influenza Hemagglutinin Fusion Domain Is an Amphipathic Helical Hairpin That Functions by Inducing Membrane Curvature. Journal of Biological Chemistry 2015, 290
(1)
, 228-238. https://doi.org/10.1074/jbc.M114.611657
- Jonathan K. Williams, Mei Hong. Probing membrane protein structure using water polarization transfer solid-state NMR. Journal of Magnetic Resonance 2014, 247 , 118-127. https://doi.org/10.1016/j.jmr.2014.08.007
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. 2D 13C–13C correlation spectra of PIV5 FPK4 in gel-phase POPC/POPG (4:1) bilayers. Shown at the top is the amino acid sequence with labeled residues color-coded according to samples. (a) GVTAA-FPK4 spectrum, measured at 253 K with 20 ms mixing. (b) VLAAT-FPK4 spectrum, coadded from two spectra measured at 243 K with 20 ms mixing and 253 K with 60 ms mixing. (c) AAQV-FPK4 spectrum, measured at 253 K with 20 ms mixing. All residues show resolved and α-helical chemical shifts. Superscript h denotes helical chemical shifts.
Figure 2
Figure 2. 2D 15N–13C correlation spectra of PIV5 FPK4 in gel-phase POPC/POPG (magenta), DOPC/DOPG (black), and POPC (blue) membranes. (a) GVTAA-FPK4 spectra. (b) VLAAT-FPK4 spectra. (c) IGALV-FPK4 spectra. (d) GVAL-FPK4 spectra. (e) AAQV-FPK4 spectra. The peptide shows predominantly β-strand chemical shifts in the POPC membrane, α-helical chemical shifts in the POPC/POPG membrane, and mixed strand and helix chemical shifts in the DOPC/DOPG membrane. Most residues in the GVTAA and VLAAT samples show two sets of chemical shifts in the DOPC/DOPG bilayer. The AAQV sample shows nearly identical α-helical chemical shifts in the POPC/POPG and DOPC/DOPG membranes. Superscripts h and s denote helical and strand chemical shifts, respectively.
Figure 3
Figure 3. 13C and 15N secondary chemical shifts of FPK4 in (a) POPC/POPG and (b) DOPC/DOPG membranes. FPK4 shows clear α-helical chemical shifts (red) in POPC/POPG bilayers and mixed helical and strand chemical shifts (blue) in DOPC/DOPG bilayers. The random coil values of Zhang et al. (86) were used to calculate the secondary shifts.
Figure 4
Figure 4. Depth of insertion of FPK4 in the POPC/POPG membrane from gel-phase spin diffusion. (a) Representative 2D spectra with 0 and 25 ms spin diffusion mixing at 258 K. (b) 1H cross sections for the peptide Cα peaks (red) and the lipid CH2 peak (black). Already at 4 ms, the peptide and lipid 1H cross sections have similar intensity patterns, indicating that the peptide is well inserted into the membrane. (c) 13C cross sections extracted from the water (blue) and lipid CH2 (black) 1H chemical shifts from the 4 ms 2D spectra. The N- and C-terminal residues have higher water/lipid intensity ratios than the middle residues. (d) Water/lipid intensity ratios for all labeled sites.
Figure 5
Figure 5. Representative 1D 13C CP MAS spectra of DOPC/DOPG-bound FPK4 as a function of temperature. The VLAAT-FPK4 spectra are shown. At high temperature, mainly β-strand chemical shifts (blue dotted lines) are observed, while at low temperature, both α-helical (red dashed lines) and β-strand chemical shifts are detected.
Figure 6
Figure 6. 2D 13C–13C correlation spectra of DOPC/DOPG-bound FPK4 in the gel phase (233 or 243 K, left column) and the LC phase (303 K, right column). (a, b) GVTAA-FPK4 spectra. (c, d) VLAAT-FPK4 spectra. (e, f) IGALV-FPK4 spectra. (g, h) AAQV-FPK4 spectra. At both temperatures, many residues show mixed α-helical and β-strand chemical shifts.
Figure 7
Figure 7. Depth of insertion of FPK4 in the DOPC/DOPG membrane from gel-phase spin diffusion spectra measured at 243 K. (a) 1H cross sections of the peptide Cα peaks (red) and lipid CH2 peak (black). By 4 ms, the peptide and lipid signals have equilibrated, indicating that the peptide is well inserted into the membrane. (b) 13C cross sections from the water (blue) and lipid CH2 (black) 1H chemical shifts of the 4 ms 2D spectra. The C-terminal α-helical residues have higher water cross peaks than the N-terminal β-strand residues, and the α-helical A123/A124 have higher water cross peaks than the β-strand A123/A124. (c) Water/lipid intensity ratios of all labeled residues in the DOPC/DOPG membrane (blue and red symbols). The β-strand residues have lower water exposure than the α-helical residues. Open symbols indicate the minor conformation. For comparison, the POPC/POPG-bound FPK4 data are also shown (black open symbols).
Figure 8
Figure 8. FPK4 interaction with the DOPE membrane. (a, b) Static 31P spectra of the membrane without (a) and with (b) FPK4 from 273 to 313 K. FPK4 increased the Lα-to-HII phase transition temperature and caused a small isotropic peak. (c) 2D 31P–1H correlation spectrum of FPK4-bound DOPE membrane with a spin diffusion mixing time of 225 ms. (d) 1H cross sections from the 2D 31P–1H spectra of peptide-free and peptide-bound DOPE membranes, compared with the 1D 1H single-pulse spectrum (top). The FPK4-bound DOPE membrane has a much weaker water–31P cross peak than the peptide-free membrane.
Figure 9
Figure 9. Conformation and depth of FPK4 in the DOPE membrane. (a) 2D 13C–13C correlation spectrum of a fresh GVTAA-FPK4 sample at 243 K. The peptide exhibits both helix and strand signals. (b) 13C CP-MAS spectra of the initial and equilibrated GVTAA-FPK4 at 246 K. At equilibrium, most residues exhibit β-strand chemical shifts. (c) 100 ms 2D 13C–1H correlation spectrum at 293 K, in the HII phase membrane. Lipid–peptide cross peaks are observed, indicating that the β-strand peptide is inserted into the hydrophobic region of the DOPE membrane.
Figure 10
Figure 10. PIV5 fusion peptide conformations in lipid membranes from solid-state NMR and outside the membrane from crystal structures. (a) Fusion peptide is fully α-helical in POPC/POPG bilayers but adopts a mixed strand/helix conformation in DOPC/DOPG bilayers. The peptide is inserted into both membranes, but the depicted tilt angle is hypothetical. The structures were built using (φ, ψ) torsion angles predicted by TALOS+. (b) Prefusion crystal structures of the PIV5 F protein in the uncleaved (green) (22) and cleaved (red) (21) states. The fusion peptide domain has similar conformations before and after cleavage and has a bend near T117. (c) Prefusion crystal structures of the influenza HA in the uncleaved (green) (76) and cleaved (red) (77) states. The N-terminal half of the fusion peptide is rotated around N12 before and after cleavage. (d) Postfusion crystal structure of the PIV5 F HRA/HRB complex. (14) Seven residues (T122–V128) of the fusion peptide are detected and show α-helical structure extended from HRA. (e) Schematic of the PIV5 fusion peptide conformation in the DOPE membrane. The lipid cylinders and water radius are drawn to scale using 15–18 water molecules per lipid based on the DOPE phase diagram. (65, 66) (f) The hemifusion stalk intermediate showing both negative and positive membrane curvatures and dehydration between two opposing bilayers. Dashed lines indicate the middle of two lipid leaflets.
References
This article references 86 other publications.
- 1Fields, C. G.; Lloyd, D. H.; Macdonald, R. L.; Ottenson, K. M.; Nobel, R. L. Peptide Res. 1991, 4, 951HBTU activation for automated Fmoc solid-phase peptide synthesisFields, C. G.; Lloyd, D. H.; Macdonald, R. L.; Otteson, K. M.; Noble, Richard L.Peptide Research (1991), 4 (2), 95-101CODEN: PEREEO; ISSN:1040-5704.Excellent results have been obtained for the 9-fluorenylmethoxycarbonyl (Fmoc) solid-phase syntheses of peptides using the activating reagent 2-(1H-benzotriazol-1-yl)-1,1,3,3-tetramethyluronium hexafluorophosphate (HBTU). Activation occurs very rapidly in DMF and N-methylpyrrolidone, optimal solvents for peptide resin solvation. Complete coupling reactions occur in only 10-30 min. Residues such as arginine, isoleucine, leucine, and valine, which often require double coupling by other activation methods, react with high efficiency by single coupling when HBTU is used. The Fmoc/HBTU chem. has recently been applied to the peptide synthesizers. The incorporation of trityl side-chain protection for Fmoc-Asn and Fmoc-Gln further enhances coupling efficiencies in difficult sequences.
- 2Chang, A.; Dutch, R. E. Viruses 2012, 4, 6132Paramyxovirus fusion and entry: multiple paths to a common endChang, Andres; Dutch, Rebecca E.Viruses (2012), 4 (), 613-636CODEN: VIRUBR; ISSN:1999-4915. (MDPI AG)A review. The paramyxovirus family contains many common human pathogenic viruses, including measles, mumps, the parainfluenza viruses, respiratory syncytial virus, human metapneumovirus, and the zoonotic henipaviruses, Hendra and Nipah. While the expression of a type 1 fusion protein and a type 2 attachment protein is common to all paramyxoviruses, there is considerable variation in viral attachment, the activation and triggering of the fusion protein, and the process of viral entry. In this review, we discuss recent advances in the understanding of paramyxovirus F protein-mediated membrane fusion, an essential process in viral infectivity. We also review the role of the other surface glycoproteins in receptor binding and viral entry, and the implications for viral infection. Throughout, we conc. on the commonalities and differences in fusion triggering and viral entry among the members of the family. Finally, we highlight key unanswered questions and how further studies can identify novel targets for the development of therapeutic treatments against these human pathogens.
- 3Dutch, R. E. PLoS Pathog. 2010, 6, e1000881There is no corresponding record for this reference.
- 4Lamb, R. A.; Jardetzky, T. S. Curr. Opin. Struct. Biol. 2007, 17, 4274Structural basis of viral invasion: lessons from paramyxovirus FLamb, Robert A.; Jardetzky, Theodore S.Current Opinion in Structural Biology (2007), 17 (4), 427-436CODEN: COSBEF; ISSN:0959-440X. (Elsevier B.V.)A review. The structures of glycoproteins that mediate enveloped virus entry into cells have revealed dramatic structural changes that accompany membrane fusion and provided mechanistic insights into this process. The group of class I viral fusion proteins includes the influenza hemagglutinin, paramyxovirus F, HIV env, and other mechanistically related fusogens, but these proteins are unrelated in sequence and exhibit clearly distinct structural features. Recently detd. crystal structures of the paramyxovirus F protein in two conformations, representing pre-fusion and post-fusion states, reveal a novel protein architecture that undergoes large-scale, irreversible refolding during membrane fusion, extending our understanding of this diverse group of membrane fusion machines.
- 5Russell, C. J.; Luque, L. E. Trends Microbiol. 2006, 14, 243There is no corresponding record for this reference.
- 6Horvath, C. M.; Lamb, R. A. J. Virol. 1992, 66, 24436Studies on the fusion peptide of a paramyxovirus fusion glycoprotein: roles of conserved residues in cell fusionHorvath, Curt M.; Lamb, Robert A.Journal of Virology (1992), 66 (4), 2443-55CODEN: JOVIAM; ISSN:0022-538X.The role of residues in the conserved hydrophobic N-terminal fusion peptide of the paramyxovirus fusion (F) protein in causing cell-cell fusion was examd. Mutations were introduced into the cDNA encoding the simian virus 5 (SV5) F protein, the altered F proteins were expressed by using an eukaryotic vector, and their ability to mediate syncytium formation was detd. The mutant F proteins contained both single- and multiple-amino-acid substitutions, and they exhibited a variety of intracellular transport properties and fusion phenotypes. The data indicate that many substitutions in the conserved amino acids of the simian virus 5 F fusion peptide can be tolerated without loss of biol. activity. Mutant F proteins which were not transported to the cell surface did not cause cell-cell fusion, but all of the mutants which were transported to the cell surface were fusion competent, exhibiting fusion properties similar to or better than those of the wild-type F protein. Mutant F proteins contg. glycine-to-alanine substitutions had altered intracellular transport characteristics, yet they exhibited a great increase in fusion activity. The potential structural implications of this substitution and the possible importance of these glycine residues in maintaining appropriate levels of fusion activity are discussed.
- 7Russell, C. J.; Jardetzky, T. S.; Lamb, R. A. J. Virol. 2004, 78, 137277Conserved glycine residues in the fusion peptide of the paramyxovirus fusion protein regulate activation of the native stateRussell, Charles J.; Jardetzky, Theodore S.; Lamb, Robert A.Journal of Virology (2004), 78 (24), 13727-13742CODEN: JOVIAM; ISSN:0022-538X. (American Society for Microbiology)Hydrophobic fusion peptides (FPs) are the most highly conserved regions of class I viral fusion-mediating glycoproteins (vFGPs). FPs often contain conserved glycine residues thought to be crit. for forming structures that destabilize target membranes. Unexpectedly, a mutation of glycine residues in the FP of the fusion (F) protein from the paramyxovirus simian parainfluenza virus 5 (SV5) resulted in mutant F proteins with hyperactive fusion phenotypes (C. M. Horvath and R. A. Lamb, J. Virol.66:2443-2455, 1992). Here, we constructed G3A and G7A mutations into the F proteins of SV5 (W3A and WR isolates), Newcastle disease virus (NDV), and human parainfluenza virus type 3 (HPIV3). All of the mutant F proteins, except NDV G7A, caused increased cell-cell fusion despite having slight to moderate redns. in cell surface expression compared to those of wild-type F proteins. The G3A and G7A mutations cause SV5 WR F, but not NDV F or HPIV3 F, to be triggered to cause fusion in the absence of coexpression of its homo-typic receptor-binding protein hemagglutinin-neuraminidase (HN), suggesting that NDV and HPIV3 F have stricter requirements for homotypic HN for fusion activation. Dye transfer assays show that the G3A and G7A mutations decrease the energy required to activate F at a step in the fusion cascade preceding prehairpin intermediate formation and hemifusion. Conserved glycine residues in the FP of paramyxovirus F appear to have a primary role in regulating the activation of the metastable native form of F. Glycine residues in the FPs of other class I vFGPs may also regulate fusion activation.
- 8Bissonnette, M. L.; Donald, J. E.; DeGrado, W. F.; Jardetzky, T. S.; Lamb, R. A. J. Mol. Biol. 2009, 386, 14There is no corresponding record for this reference.
- 9Baquero, E.; Albertini, A. A.; Vachette, P.; Lepault, J.; Bressanelli, S.; Gaudin, Y. Curr. Opin. Virol. 2013, 3, 1439Intermediate conformations during viral fusion glycoprotein structural transitionBaquero, Eduard; Albertini, Aurelie A.; Vachette, Patrice; Lepault, Jean; Bressanelli, Stephane; Gaudin, YvesCurrent Opinion in Virology (2013), 3 (2), 143-150CODEN: COVUAF; ISSN:1879-6257. (Elsevier B. V.)A review. Entry of enveloped viruses into cells requires the fusion of viral and cellular membranes, driven by conformational changes in viral glycoproteins. Three different classes of viral fusion proteins have been hitherto identified based on common structural elements. Crystal structures have provided static pictures of pre-fusion and post-fusion conformations of these proteins and have revealed the dramatic reorganization of the mols., but the transition pathway remains elusive. In this review, we will focus on recent data aiming to characterize intermediate structures during the conformational change. All these data support the existence of a pre-hairpin intermediate, but its oligomeric status is still a matter of debate.
- 10Weissenhorn, W.; Hinz, A.; Gaudin, Y. FEBS Lett. 2007, 581, 2150There is no corresponding record for this reference.
- 11White, J. M.; Delos, S. E.; Brecher, M.; Schornberg, K. Crit. Rev. Biochem. Mol. Biol. 2008, 43, 18911Structures and Mechanisms of Viral Membrane Fusion Proteins: Multiple Variations on a Common ThemeWhite, Judith M.; Delos, Sue E.; Brecher, Matthew; Schornberg, KathrynCritical Reviews in Biochemistry and Molecular Biology (2008), 43 (3), 189-219CODEN: CRBBEJ; ISSN:1040-9238. (Informa Healthcare)A review. Recent work has identified three distinct classes of viral membrane fusion proteins based on structural criteria. In addn., there are at least four distinct mechanisms by which viral fusion proteins can be triggered to undergo fusion-inducing conformational changes. Viral fusion proteins also contain different types of fusion peptides and vary in their reliance on accessory proteins. These differing features combine to yield a rich diversity of fusion proteins. Yet despite this staggering diversity, all characterized viral fusion proteins convert from a fusion-competent state (dimers or trimers, depending on the class) to a membrane-embedded homotrimeric prehairpin, and then to a trimer-of-hairpins that brings the fusion peptide, attached to the target membrane, and the transmembrane domain, attached to the viral membrane, into close proximity thereby facilitating the union of viral and target membranes. During these conformational conversions, the fusion proteins induce membranes to progress through stages of close apposition, hemifusion, and then the formation of small, and finally large, fusion pores. Clearly, highly divergent proteins have converged on the same overall strategy to mediate fusion, an essential step in the life cycle of every enveloped virus.
- 12Lamb, R. A.; Paterson, R. G.; Jardetzky, T. S. Virology 2006, 344, 30There is no corresponding record for this reference.
- 13Harrison, S. C. Nat. Struct. Mol. Biol. 2008, 15, 69013Viral membrane fusionHarrison, Stephen C.Nature Structural & Molecular Biology (2008), 15 (7), 690-698CODEN: NSMBCU; ISSN:1545-9993. (Nature Publishing Group)A review. Infection by viruses having lipid-bilayer envelopes proceeds through fusion of the viral membrane with a membrane of the target cell. Viral 'fusion proteins' facilitate this process. They vary greatly in structure, but all seem to have a common mechanism of action, in which a ligand-triggered, large-scale conformational change in the fusion protein is coupled to apposition and merger of the 2 bilayers. The authors describe 3 examples: (1) the influenza virus hemagglutinin; (2) the flavivirus E protein; and (3) the vesicular stomatitis virus G protein, in some detail in order to illustrate the ways in which different structures have evolved to implement this common mechanism. Fusion inhibitors can be effective antiviral agents.
- 14Baker, K. A.; Dutch, R. E.; Lamb, R. A.; Jardetzky, T. S. Mol. Cell 1999, 3, 309There is no corresponding record for this reference.
- 15Chan, D. C.; Kim, P. S. Cell 1998, 93, 681There is no corresponding record for this reference.
- 16McLellan, J. S.; Yang, Y.; Graham, B. S.; Kwong, P. D. J. Virol. 2011, 85, 778816Structure of respiratory syncytial virus fusion glycoprotein in the postfusion conformation reveals preservation of neutralizing epitopesMcLellan, Jason S.; Yang, Yongping; Graham, Barney S.; Kwong, Peter D.Journal of Virology (2011), 85 (15), 7788-7796CODEN: JOVIAM; ISSN:0022-538X. (American Society for Microbiology)Respiratory syncytial virus (RSV) invades host cells via a type I fusion (F) glycoprotein that undergoes dramatic structural rearrangements during the fusion process. Neutralizing monoclonal antibodies, such as 101F, palivizumab, and motavizumab, target two major antigenic sites on the RSV F glycoprotein. The structures of these sites as peptide complexes with motavizumab and 101F have been previously detd., but a structure for the trimeric RSV F glycoprotein ectodomain has remained elusive. To address this issue, we undertook structural and biophys. studies on stable ectodomain constructs. Here, we present the 2.8-Å crystal structure of the trimeric RSV F ectodomain in its postfusion conformation. The structure revealed that the 101F and motavizumab epitopes are present in the postfusion state and that their conformations are similar to those obsd. in the antibody-bound peptide structures. Both antibodies bound the postfusion F glycoprotein with high affinity in surface plasmon resonance expts. Modeling of the antibodies bound to the F glycoprotein predicts that the 101F epitope is larger than the linear peptide and restricted to a single protomer in the trimer, whereas motavizumab likely contacts residues on two protomers, indicating a quaternary epitope. Mechanistically, these results suggest that 101F and motavizumab can bind to multiple conformations of the fusion glycoprotein and can neutralize late in the entry process. The structural preservation of neutralizing epitopes in the postfusion state suggests that this conformation can elicit neutralizing antibodies and serve as a useful vaccine antigen.
- 17Swanson, K.; Wen, X.; Leser, G. P.; Paterson, R. G.; Lamb, R. A.; Jardetzky, T. S. Virology 2010, 402, 372There is no corresponding record for this reference.
- 18Tan, K.; Liu, J.; Wang, J.; Shen, S.; Lu, M. Proc. Natl. Acad. Sci. U.S.A. 1997, 94, 12303There is no corresponding record for this reference.
- 19Yin, H. S.; Paterson, R. G.; Wen, X.; Lamb, R. A.; Jardetzky, T. S. Proc. Natl. Acad. Sci. U.S.A. 2005, 102, 9288There is no corresponding record for this reference.
- 20Zhao, X.; Singh, M.; Malashkevich, V. N.; Kim, P. S. Proc. Natl. Acad. Sci. U.S.A. 2000, 97, 14172There is no corresponding record for this reference.
- 21Welch, B. D.; Liu, Y.; Kors, C. A.; Leser, G. P.; Jardetzky, T. S.; Lamb, R. A. Proc. Natl. Acad. Sci. U.S.A. 2012, 109, 16672There is no corresponding record for this reference.
- 22Yin, H. S.; Wen, X.; Paterson, R. G.; Lamb, R. A.; Jardetzky, T. S. Nature 2006, 439, 38There is no corresponding record for this reference.
- 23Kim, Y. H.; Donald, J. E.; Grigoryan, G.; Leser, G. P.; Fadeev, A. Y.; Lamb, R. A.; DeGrado, W. F. Proc. Natl. Acad. Sci. U.S.A. 2011, 108, 20992There is no corresponding record for this reference.
- 24Han, X.; Bushweller, J. H.; Cafiso, D. S.; Tamm, L. K. Nat. Struct. Biol. 2001, 8, 715There is no corresponding record for this reference.
- 25Jaroniec, C. P.; Kaufman, J. D.; Stahl, S. J.; Viard, M.; Blumenthal, R.; Wingfield, P. T.; Bax, A. Biochemistry 2005, 44, 16167There is no corresponding record for this reference.
- 26Lai, A. L.; Tamm, L. K. J. Biol. Chem. 2010, 285, 37467There is no corresponding record for this reference.
- 27Li, Y.; Han, X.; Lai, A. L.; Bushweller, J. H.; Cafiso, D. S.; Tamm, L. K. J. Virol. 2005, 79, 1206527Membrane structures of the hemifusion-inducing fusion peptide: mutant G1S and the fusion-blocking mutant G1V of influenza virus hemagglutinin suggest a mechanism for pore opening in membrane fusionLi, Yinling; Han, Xing; Lai, Alex L.; Bushweller, John H.; Cafiso, David S.; Tamm, Lukas K.Journal of Virology (2005), 79 (18), 12065-12076CODEN: JOVIAM; ISSN:0022-538X. (American Society for Microbiology)Influenza virus hemagglutinin (HA)-mediated membrane fusion is initiated by a conformational change that releases a V-shaped hydrophobic fusion domain, the fusion peptide, into the lipid bilayer of the target membrane. The most N-terminal residue of this domain, a glycine, is highly conserved and is particularly crit. for HA function; G1S and G1V mutant HAs cause hemifusion and abolish fusion, resp. We have detd. the at. resoln. structures of the G1S and G1V mutant fusion domains in membrane environments. G1S forms a V with a disrupted "glycine edge" on its N-terminal arm and G1V adopts a slightly tilted linear helical structure in membranes. Abolishment of the kink in G1V results in reduced hydrophobic penetration of the lipid bilayer and an increased propensity to form β-structures at the membrane surface. These results underline the functional importance of the kink in the fusion peptide and suggest a structural role for the N-terminal glycine ridge in viral membrane fusion.
- 28Li, Y.; Tamm, L. K. Biophys. J. 2007, 93, 876There is no corresponding record for this reference.
- 29Lorieau, J. L.; Louis, J. M.; Bax, A. Proc. Natl. Acad. Sci. U.S.A. 2010, 107, 11341There is no corresponding record for this reference.
- 30Lorieau, J. L.; Louis, J. M.; Schwieters, C. D.; Bax, A. Proc. Natl. Acad. Sci. U.S.A. 2012, 109, 19994There is no corresponding record for this reference.
- 31Lai, A. L.; Tamm, L. K. J. Biol. Chem. 2007, 282, 23946There is no corresponding record for this reference.
- 32Tamm, L. K.; Lai, A. L.; Li, Y. Biochim. Biophys. Acta 2007, 1768, 3052There is no corresponding record for this reference.
- 33Epand, R. M. Biochim. Biophys. Acta 2003, 1614, 116There is no corresponding record for this reference.
- 34Li, Y.; Han, X.; Tamm, L. K. Biochemistry 2003, 42, 7245There is no corresponding record for this reference.
- 35Sun, Y.; Weliky, D. P. J. Am. Chem. Soc. 2009, 131, 13228There is no corresponding record for this reference.
- 36Lorieau, J. L.; Louis, J. M.; Bax, A. Biopolymers 2013, 99, 189There is no corresponding record for this reference.
- 37Gordon, L. M.; Mobley, P. W.; Lee, W.; Eskandari, S.; Kaznessis, Y. N.; Sherman, M. A.; Waring, A. J. Protein Sci. 2004, 13, 101237Conformational mapping of the N-terminal peptide of HIV-1 gp41 in lipid detergent and aqueous environments using 13C-enhanced Fourier transform infrared spectroscopyGordon, Larry M.; Mobley, Patrick W.; Lee, William; Eskandari, Sepehr; Kaznessis, Yiannis N.; Sherman, Mark A.; Waring, Alan J.Protein Science (2004), 13 (4), 1012-1030CODEN: PRCIEI; ISSN:0961-8368. (Cold Spring Harbor Laboratory Press)The N-terminal domain of HIV-1 glycoprotein 41,000 (gp41) participates in viral fusion processes. Here, we use phys. and computational methodologies to examine the secondary structure of a peptide based on the N terminus (FP; residues 1-23) in aq. and detergent environments. 12C-Fourier transform IR (FTIR) spectroscopy indicated greater α-helix for FP in lipid-detergent SDS and aq. phosphate-buffered saline (PBS) than in only PBS. 12C-FTIR spectra also showed disordered FP conformations in these two environments, along with substantial β-structure for FP alone in PBS. In expts. that map conformations to specific residues, isotope-enhanced FTIR spectroscopy was performed using FP peptides labeled with 13C-carbonyl. 13C-FTIR results on FP in SDS at low peptide loading indicated α-helix (residues 5 to 16) and disordered conformations (residues 1-4). Because earlier 13C-FTIR anal. of FP in lipid bilayers demonstrated α-helix for residues 1-16 at low peptide loading, the FP structure in SDS micelles only approximates that found for FP with membranes. Mol. dynamics simulations of FP in an explicit SDS micelle indicate that the fraying of the first three to four residues may be due to the FP helix moving to one end of the micelle. In PBS alone, however, electron microscopy of FP showed large fibrils, while 13C-FTIR spectra demonstrated antiparallel β-sheet for FP (residues 1-12), analogous to that reported for amyloid peptides. Because FP and amyloid peptides each exhibit plaque formation, α-helix to β-sheet interconversion, and membrane fusion activity, amyloid and N-terminal gp41 peptides may belong to the same superfamily of proteins.
- 38Qiang, W.; Bodner, M. L.; Weliky, D. P. J. Am. Chem. Soc. 2008, 130, 5459There is no corresponding record for this reference.
- 39Qiang, W.; Sun, Y.; Weliky, D. P. Proc. Natl. Acad. Sci. U.S.A. 2009, 106, 15314There is no corresponding record for this reference.
- 40Lai, A. L.; Moorthy, A. E.; Li, Y.; Tamm, L. K. J. Mol. Biol. 2012, 418, 3There is no corresponding record for this reference.
- 41Gordon, L. M.; Mobley, P. W.; Pilpa, R.; Sherman, M. A.; Waring, A. J. Biochim. Biophys. Acta 2002, 1559, 96There is no corresponding record for this reference.
- 42Rafalski, M.; Lear, J. D.; DeGrado, W. F. Biochemistry 1990, 29, 7917There is no corresponding record for this reference.
- 43Saez-Cirion, A.; Nieva, J. L. Biochim. Biophys. Acta 2002, 1564, 57There is no corresponding record for this reference.
- 44Gabrys, C. M.; Qiang, W.; Sun, Y.; Xie, L.; Schmick, S. D.; Weliky, D. P. J. Phys. Chem. A 2013, 117, 9848There is no corresponding record for this reference.
- 45Hong, M.; Zhang, Y.; Hu, F. Annu. Rev. Phys. Chem. 2012, 63, 1There is no corresponding record for this reference.
- 46Fuhrmans, M.; Marrink, S. J. J. Am. Chem. Soc. 2012, 134, 1543There is no corresponding record for this reference.
- 47Siegel, D. P. Biophys. J. 1999, 76, 291There is no corresponding record for this reference.
- 48Kasson, P. M.; Pande, V. S. PLoS Comput. Biol. 2007, 3, e220There is no corresponding record for this reference.
- 49Chernomordik, L. V.; Kozlov, M. M. Cell 2005, 123, 375There is no corresponding record for this reference.
- 50Tamm, L. K.; Han, X. Biosci. Rep. 2000, 20, 50150Viral fusion peptides: a tool set to disrupt and connect biological membranesTamm, Lukas K.; Han, XingBioscience Reports (2000), 20 (6), 501-518CODEN: BRPTDT; ISSN:0144-8463. (Kluwer Academic/Plenum Publishers)A review, with 56 refs., on the structure and function of viral fusion peptides. The fusion peptides of influenza virus hemagglutinin and human immunodeficiency virus are used as paradigms. Fusion peptides assocd. with lipid bilayers are conformationally polymorphic. Current evidence suggests that the fusion-promoting state is the obliquely inserted α-helix. Fusion peptides also have a tendency to self-assoc. into β-sheets at membrane surfaces. Although the conformational conversion between α- and β-states is reversible under controlled conditions, its physiol. relevance is not yet known. The energetics of peptide insertion and self-assocn. could be measured recently using more sol. 2nd generation fusion peptides. Fusion peptides have been reported to change membrane curvature and the state of hydration of membrane surfaces. The combined results are built into a model for the mechanism by which fusion peptides are proposed to assist in biol. membrane fusion.
- 51Siegel, D. P.; Epand, R. M. Biochim. Biophys. Acta 2000, 1468, 87There is no corresponding record for this reference.
- 52Yao, H.; Hong, M. J. Mol. Biol. 2013, 425, 563There is no corresponding record for this reference.
- 53Takegoshi, K.; Nakamura, S.; Terao, T. Chem. Phys. Lett. 2001, 344, 631There is no corresponding record for this reference.
- 54Hong, M.; Griffin, R. G. J. Am. Chem. Soc. 1998, 120, 7113There is no corresponding record for this reference.
- 55Huster, D.; Yao, X. L.; Hong, M. J. Am. Chem. Soc. 2002, 124, 874There is no corresponding record for this reference.
- 56Mani, R.; Cady, S. D.; Tang, M.; Waring, A. J.; Lehrer, R. I.; Hong, M. Proc. Natl. Acad. Sci. U.S.A. 2006, 103, 16242There is no corresponding record for this reference.
- 57Wang, T.; Yao, H.; Hong, M. J. Biomol. NMR 2013, 56, 139There is no corresponding record for this reference.
- 58Donald, J. E.; Zhang, Y.; Fiorin, G.; Carnevale, V.; Slochower, D. R.; Gai, F.; Klein, M. L.; DeGrado, W. F. Proc. Natl. Acad. Sci. U.S.A. 2011, 108, 3958There is no corresponding record for this reference.
- 59Shen, Y.; Delaglio, F.; Cornilescu, G.; Bax, A. J. Biomol. NMR 2009, 44, 213There is no corresponding record for this reference.
- 60Huster, D.; Yao, X.; Hong, M. J. Am. Chem. Soc. 2002, 124, 874There is no corresponding record for this reference.
- 61Cady, S. D.; Goodman, C.; Tatko, C.; DeGrado, W. F.; Hong, M. J. Am. Chem. Soc. 2007, 129, 5719There is no corresponding record for this reference.
- 62Hong, M.; Doherty, T. Chem. Phys. Lett. 2006, 432, 296There is no corresponding record for this reference.
- 63Drechsler, A.; Anderluh, G.; Norton, R. S.; Separovic, F. Biochim. Biophys. Acta 2010, 1798, 244There is no corresponding record for this reference.
- 64Traikia, M.; Warschawski, D. E.; Recouvreur, M.; Cartaud, J.; Devaux, P. F. Eur. Biophys. J. 2000, 29, 184There is no corresponding record for this reference.
- 65Gawrisch, K.; Parsegian, V. A.; Hajduk, D. A.; Tate, M. W.; Graner, S. M.; Fuller, N. L.; Rand, R. P. Biochemistry 1992, 31, 2856There is no corresponding record for this reference.
- 66Rand, R. P.; Fuller, N. L. Biophys. J. 1994, 66, 2127There is no corresponding record for this reference.
- 67Thayer, A. M.; Kohler, S. J. Biochemistry 1981, 20, 6831There is no corresponding record for this reference.
- 68Tenchov, B. G.; MacDonald, R. C.; Lentz, B. R. Biophys. J. 2013, 104, 1029There is no corresponding record for this reference.
- 69Schmidt, N.; Mishra, A.; Lai, G. H.; Wong, G. C. FEBS Lett. 2010, 584, 1806There is no corresponding record for this reference.
- 70Mishra, A.; Gordon, V. D.; Yang, L.; Coridan, R.; Wong, G. C. L. Angew. Chem., Int. Ed. 2008, 47, 2986There is no corresponding record for this reference.
- 71Kučerka, N.; Nieh, M. P.; Katsaras, J. Biochim. Biophys. Acta 2011, 1808, 2761There is no corresponding record for this reference.
- 72Kucerka, N.; Tristram-Nagle, S.; Nagle, J. F. J. Membr. Biol. 2005, 208, 19372Structure of Fully Hydrated Fluid Phase Lipid Bilayers with Monounsaturated ChainsKucerka, Norbert; Tristram-Nagle, Stephanie; Nagle, John F.Journal of Membrane Biology (2006), 208 (3), 193-202CODEN: JMBBBO; ISSN:0022-2631. (Springer)Quant. structures are obtained at 30°C for the fully hydrated fluid phases of palmitoyloleoylphosphatidylcholine (POPC), with a double bond on the sn-2 hydrocarbon chain, and for dierucoylphosphatidylcholine (di22:1PC), with a double bond on each hydrocarbon chain. The form factors F(qz) for both lipids are obtained using a combination of 3 methods: (1) Volumetric measurements provide F(0), (2) x-ray scattering from extruded unilamellar vesicles provides |F(qz)| for low qz, (3) Diffuse x-ray scattering from oriented stacks of bilayers provides |F(qz)| for high qz. Also, data using method (2) are added to the authors' recent data for dioleoylphosphatidylcholine (DOPC) using methods (1) and (3); the new DOPC data agree very well with the recent data and with (4) the authors' older data obtained using a liq. crystallog. x-ray method. The authors used hybrid electron d. models to obtain structural results from these form factors. The result for area per lipid (A) for DOPC 72.4 ± 0.5 Å2 agrees well with the authors' earlier publications, and the authors find A = 69.3 ± 0.5 Å2 for di22:1PC and A = 68.3 ± 1.5 Å2 for POPC. The authors obtain the values for 5 different av. thicknesses: hydrophobic, steric, head-head, phosphate-phosphate and Luzzati. Comparison of the results for these 3 lipids and for the authors' recent dimyristoylphosphatidylcholine (DMPC) detn. provides quant. measures of the effect of unsatn. on bilayer structure. The authors' results suggest that lipids with one monounsatd. chain have quant. bilayer structures closer to lipids with 2 monounsatd. chains than to lipids with 2 completely satd. chains.
- 73Chernomordik, L. V.; Kozlov, M. M. Annu. Rev. Biochem. 2003, 72, 17573Protein-lipid interplay in fusion and fission of biological membranesChernomordik, Leonid V.; Kozlov, Michael M.Annual Review of Biochemistry (2003), 72 (), 175-207CODEN: ARBOAW; ISSN:0066-4154. (Annual Reviews Inc.)A review. Disparate biol. processes involve fusion of two membranes into one and fission of one membrane into two. To formulate the possible job description for the proteins that mediate remodeling of biol. membranes, we analyze the energy price of disruption and bending of membrane lipid bilayers at the different stages of bilayer fusion. The phenomenol. and the pathways of the well-characterized reactions of biol. remodeling, such as fusion mediated by influenza hemagglutinin, are compared with those studied for protein-free bilayers. We briefly consider some proteins involved in fusion and fission, and the dependence of remodeling on the lipid compn. of the membranes. The specific hypothetical mechanisms by which the proteins can lower the energy price of the bilayer rearrangement are discussed in light of the exptl. data and the requirements imposed by the elastic properties of the bilayer.
- 74Markvoort, A. J.; Marrink, S. J. Curr. Top. Membr. 2011, 68, 25974Lipid acrobatics in the membrane fusion arenaMarkvoort, Albert J.; Marrink, Siewert J.Current Topics in Membranes (2011), 68 (Membrane Fusion), 259-294CODEN: CTMEET; ISSN:1063-5823. (Elsevier Inc.)A review on the recent contribution of computer simulation approaches to unravel the mol. details of membrane fusion. It considers particle-based simulations of apposed bilayers, bilayers and vesicles, or between vesicles. It sets out with a short overview of the historical background of fusion-related simulation studies, introducing the main particle-based simulation techniques used. It proceeds with a detailed description of the various fusion pathways that are obsd. in these simulations, and the energetics involved. Special attention is devoted to discuss the accumulating evidence that the first barrier to fusion is in fact the splaying of a single lipid tail. Subsequently, the process of vesicle fission is described, showing that it is not just fusion reversed. Finally, the growing body of simulations probing the effect of fusion peptides and proteins on the fusion process is discussed.
- 75Chernomordik, L. V.; Leikina, E.; Frolov, V.; Bronk, P.; Zimmerberg, J. J. Cell Biol. 1997, 136, 81There is no corresponding record for this reference.
- 76Chen, J.; Lee, K. H.; Steinhauer, D. A.; Stevens, D. J.; Skehel, J. J.; Wiley, D. C. Cell 1998, 95, 409There is no corresponding record for this reference.
- 77Weis, W. I.; Brunger, A. T.; Skehel, J. J.; Wiley, D. C. J. Mol. Biol. 1990, 212, 737There is no corresponding record for this reference.
- 78Smith, E. C.; Gregory, S. M.; Tamm, L. K.; Creamer, T. P.; Dutch, R. E. J. Biol. Chem. 2012, 287, 30035There is no corresponding record for this reference.
- 79Epand, R. M.; Epand, R. F. Biochem. Biophys. Res. Commun. 1994, 202, 1420There is no corresponding record for this reference.
- 80Epand, R. M.; Epand, R. F.; Martin, I.; Ruysschaert, J. M. Biochemistry 2001, 40, 8800There is no corresponding record for this reference.
- 81Gabrys, C. M.; Yang, R.; Wasniewski, C. M.; Yang, J.; Canlas, C. G.; Qiang, W.; Sun, Y.; Weliky, D. P. Biochim. Biophys. Acta 2010, 1798, 194There is no corresponding record for this reference.
- 82Pereira, F. B.; Valpuesta, J. M.; Basañez, G.; Goñi, F. M.; Nieva, J. L. Chem. Phys. Lipids 1999, 103, 11There is no corresponding record for this reference.
- 83Tristram-Nagle, S.; Chan, R.; Kooijman, E.; Uppamoochikkal, P.; Qiang, W.; Weliky, D. P.; Nagle, J. F. J. Mol. Biol. 2010, 402, 139There is no corresponding record for this reference.
- 84Kaasgaard, T.; Drummond, C. J. Phys. Chem. Chem. Phys. 2006, 8, 4957There is no corresponding record for this reference.
- 85Kulkarni, C. V.; Wachter, W.; Iglesias-Salto, G.; Engelskirchen, S.; Ahualli, S. Phys. Chem. Chem. Phys. 2011, 13, 3004There is no corresponding record for this reference.
- 86Zhang, H.; Neal, S.; Wishart, D. S. J. Biomol. NMR 2003, 25, 173There is no corresponding record for this reference.
Supporting Information
Supporting Information
Additional 2D spectra and a table of residue-specific helicity of FPK4. This material is available free of charge via the Internet at http://pubs.acs.org.
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.