Coralie Alonso, Alan Waring, Joseph A. Zasadzinski  Biophysical Journal 

Slides:



Advertisements
Similar presentations
Volume 93, Issue 1, Pages (July 2007)
Advertisements

Comparing Experimental and Simulated Pressure-Area Isotherms for DPPC
Effect of Microvillus Deformability on Leukocyte Adhesion Explored Using Adhesive Dynamics Simulations  Kelly E. Caputo, Daniel A. Hammer  Biophysical.
Volume 80, Issue 4, Pages (April 2001)
Antonio Cruz, Luis Vázquez, Marisela Vélez, Jesús Pérez-Gil 
Volume 96, Issue 12, Pages (June 2009)
Liquid-Crystalline Collapse of Pulmonary Surfactant Monolayers
Visualizing the Analogy between Competitive Adsorption and Colloid Stability to Restore Lung Surfactant Function  Ian C. Shieh, Alan J. Waring, Joseph A.
Jamming and no nanosilos in pure DPPC:POPG 7:3
Insertion of Alzheimer’s Aβ40 Peptide into Lipid Monolayers
Volume 89, Issue 5, Pages (November 2005)
Volume 100, Issue 6, Pages (March 2011)
The Mechanism of Collapse of Heterogeneous Lipid Monolayers
Karen Sabatini, Juha-Pekka Mattila, Paavo K.J. Kinnunen 
Volume 96, Issue 11, Pages (June 2009)
Volume 93, Issue 6, Pages (September 2007)
Juan M. Vanegas, Maria F. Contreras, Roland Faller, Marjorie L. Longo 
Volume 101, Issue 1, Pages (July 2011)
Volume 97, Issue 2, Pages (July 2009)
Volume 87, Issue 2, Pages (August 2004)
Volume 80, Issue 5, Pages (May 2001)
Scanning Near-Field Fluorescence Resonance Energy Transfer Microscopy
Simple, Helical Peptoid Analogs of Lung Surfactant Protein B
H.M. Seeger, G. Marino, A. Alessandrini, P. Facci  Biophysical Journal 
Benjamin L. Stottrup, Sarah L. Keller  Biophysical Journal 
Following the Formation of Supported Lipid Bilayers on Mica: A Study Combining AFM, QCM-D, and Ellipsometry  Ralf P. Richter, Alain R. Brisson  Biophysical.
Cholera Toxin Assault on Lipid Monolayers Containing Ganglioside GM1
Structure of Supported Bilayers Composed of Lipopolysaccharides and Bacterial Phospholipids: Raft Formation and Implications for Bacterial Resistance 
Volume 93, Issue 2, Pages (July 2007)
Pulmonary Surfactant Protein SP-C Counteracts the Deleterious Effects of Cholesterol on the Activity of Surfactant Films under Physiologically Relevant.
Volume 74, Issue 5, Pages (May 1998)
Volume 98, Issue 9, Pages (May 2010)
Roman Volinsky, Riku Paananen, Paavo K.J. Kinnunen  Biophysical Journal 
Volume 95, Issue 6, Pages (September 2008)
T.M. Okonogi, H.M. McConnell  Biophysical Journal 
V.P. Ivanova, I.M. Makarov, T.E. Schäffer, T. Heimburg 
Sarah L. Veatch, Sarah L. Keller  Biophysical Journal 
Volume 84, Issue 1, Pages (January 2003)
Obstructed Diffusion in Phase-Separated Supported Lipid Bilayers: A Combined Atomic Force Microscopy and Fluorescence Recovery after Photobleaching Approach 
G. Garbès Putzel, Mark J. Uline, Igal Szleifer, M. Schick 
Volume 96, Issue 11, Pages (June 2009)
Scanning Force Microscopy at the Air-Water Interface of an Air Bubble Coated with Pulmonary Surfactant  D. Knebel, M. Sieber, R. Reichelt, H.-J. Galla,
Volume 113, Issue 6, Pages (September 2017)
Jason K. Cheung, Thomas M. Truskett  Biophysical Journal 
Tapani Viitala, Jouko Peltonen  Biophysical Journal 
Volume 78, Issue 1, Pages (January 2000)
Spontaneous Formation of Two-Dimensional and Three-Dimensional Cholesterol Crystals in Single Hydrated Lipid Bilayers  Roy Ziblat, Iael Fargion, Leslie.
On the Role of Acylation of Transmembrane Proteins
Multiple Folding Pathways of the SH3 Domain
Volume 101, Issue 9, Pages (November 2011)
Interaction of Lung Surfactant Proteins with Anionic Phospholipids
Philip J. Robinson, Teresa J.T. Pinheiro  Biophysical Journal 
Phospholipase D Activity Is Regulated by Product Segregation and the Structure Formation of Phosphatidic Acid within Model Membranes  Kerstin Wagner,
Lipid Asymmetry in DLPC/DSPC-Supported Lipid Bilayers: A Combined AFM and Fluorescence Microscopy Study  Wan-Chen Lin, Craig D. Blanchette, Timothy V.
Sergi Garcia-Manyes, Gerard Oncins, Fausto Sanz  Biophysical Journal 
Volume 113, Issue 6, Pages (September 2017)
Volume 94, Issue 8, Pages (April 2008)
Molecular Dynamics Simulations of Hydrophilic Pores in Lipid Bilayers
Volume 101, Issue 11, Pages (December 2011)
Juan M. Vanegas, Maria F. Contreras, Roland Faller, Marjorie L. Longo 
Alternative Mechanisms for the Interaction of the Cell-Penetrating Peptides Penetratin and the TAT Peptide with Lipid Bilayers  Semen Yesylevskyy, Siewert-Jan.
Comparing Experimental and Simulated Pressure-Area Isotherms for DPPC
Change in Rigidity in the Activated Form of the Glucose/Galactose Receptor from Escherichia coli: A Phenomenon that Will Be Key to the Development of.
Pulmonary Surfactant Model Systems Catch the Specific Interaction of an Amphiphilic Peptide with Anionic Phospholipid  Hiromichi Nakahara, Sannamu Lee,
Main Phase Transitions in Supported Lipid Single-Bilayer
Volume 79, Issue 1, Pages (July 2000)
Volume 90, Issue 9, Pages (May 2006)
Domain Growth, Shapes, and Topology in Cationic Lipid Bilayers on Mica by Fluorescence and Atomic Force Microscopy  Ariane E. McKiernan, Timothy V. Ratto,
Hong Xing You, Xiaoyang Qi, Gregory A. Grabowski, Lei Yu 
Volume 97, Issue 3, Pages (August 2009)
Presentation transcript:

Keeping Lung Surfactant Where It Belongs: Protein Regulation of Two-Dimensional Viscosity  Coralie Alonso, Alan Waring, Joseph A. Zasadzinski  Biophysical Journal  Volume 89, Issue 1, Pages 266-273 (July 2005) DOI: 10.1529/biophysj.104.052092 Copyright © 2005 The Biophysical Society Terms and Conditions

Figure 1 Surface shear viscosity, ηs, as a function of surface pressure, π, for Langmuir monolayers of POPG, PA, DPPC, DPPC/POPG 4:1, DPPC/POPG/PA 68:22:8 (wt/wt) at 25°C. The POPG monolayer remains in the fluid state at all π, thus ηs remains low. The PA monolayer undergoes a transition from a tilted to a nontilted phase at 22 mN/m, which triggers a dramatic increase of ηs. DPPC surface viscosity in the LE phase (below ∼15 mN/m) is similar to POPG. In the LC phase, ηs increases exponentially. Both DPPC/POPG (bottom inset image) and DPPC/POPG/PA (top inset image) films show coexistence between solid phase islands (black in inset images) and a continuous fluid phase (bright in inset images) for surface pressures >10 mN/m. Both images shown are taken at 30 mN/m; PA increases the fraction of solid phase at a given π, which also increases ηs. Biophysical Journal 2005 89, 266-273DOI: (10.1529/biophysj.104.052092) Copyright © 2005 The Biophysical Society Terms and Conditions

Figure 2 Surface-pressure-area isotherm of DPPC/POPG/PA (68:22:8wt %) monolayers with varying weight fractions of SP-Cff. Adding SP-Cff shifts the isotherms to larger area/molecule, showing that the protein has incorporated into the interface. The pronounced plateau at ∼40 mN/m shows that the monolayer loses material or is strongly condensed. This is consistent with the multilayer formation shown in Fig. 4. Biophysical Journal 2005 89, 266-273DOI: (10.1529/biophysj.104.052092) Copyright © 2005 The Biophysical Society Terms and Conditions

Figure 3 Surface shear viscosity, ηs, as a function of surface pressure, π, for DPPC/POPG/PA (68:22:8wt %) monolayers with varying weight fractions of SP-Cff. The surface viscosity on compression remains low up to ∼40 mN/m, at which it increases by more than three orders of magnitude. On expansion, the surface viscosity decreases more slowly with decreasing surface pressure, showing a marked hysteresis. Five percent SP-Cff increases the surface viscosity by a factor of 5 over the lipid alone. Increasing the SP-Cff to 15% provides only a marginal increase in the surface viscosity. Biophysical Journal 2005 89, 266-273DOI: (10.1529/biophysj.104.052092) Copyright © 2005 The Biophysical Society Terms and Conditions

Figure 4 AFM images of DPPC/POPG/PA+SP-Cff films. (A) Below the plateau in the isotherm, the fluid phase regions are 1nm lower (see height trace) than the condensed regions. (B) Above the plateau, the fluid regions have thickened to be ∼5nm higher than the condensed domains. This is consistent with the formation of three-layer thick areas in the fluid phase. The plateau in the isotherms corresponds to the removal of fluid phase lipids from the interface to form these multilayer patches. SP-C acts to regulate the solid/fluid phase ratio causing the surface viscosity to diverge. Biophysical Journal 2005 89, 266-273DOI: (10.1529/biophysj.104.052092) Copyright © 2005 The Biophysical Society Terms and Conditions

Figure 5 Pressure-area isotherms of 68:22:8 DPPC/POPG/PA lipid mixture with various amounts of dSP-B1-25, a peptide mimic of native SP-B (25). Increasing the dSP-B1-25 fraction moves the isotherms to the right, indicating that the dSP-B1-25 is retained in the film. The collapse pressure stays the same. Biophysical Journal 2005 89, 266-273DOI: (10.1529/biophysj.104.052092) Copyright © 2005 The Biophysical Society Terms and Conditions

Figure 6 Surface viscosity as a function of surface pressure for the dSP-B1-25 containing films. Increasing the dSP-B1-25 content to 5% only increases the surface viscosity by ∼50%, compared to more than a factor of 5 increase caused by an equal amount of SP-C. Increasing the dSP-B1-25 fraction increases the surface viscosity, suggesting that more fluid phase lipids are removed from the monolayer, consistent with the images in Fig. 7. Biophysical Journal 2005 89, 266-273DOI: (10.1529/biophysj.104.052092) Copyright © 2005 The Biophysical Society Terms and Conditions

Figure 7 AFM images of lipid films transferred to mica substrates at π=40 mN/m. (A) 0wt % dSP-B1-25. With no dSP-B1-25, the solid phase is smooth and slightly higher (∼0.5nm) than the fluid phase, which is mottled light and dark gray. (B) 5wt % dSP-B1-25. As is the case for SP-C, SP-B is located in the fluid phase domains. The solid phase domains remain smooth and are free of protrusion. dSP-B1-25 leads to the formation of isolated small protrusions from the fluid phase domains, with heights ranging from ∼5−10nm. SP-B removes fluid phase lipids from the monolayer on compression, but not nearly as efficiently as SP-C. (C) 10wt % dSP-B1-25. Increasing the dSP-B1-25 to 10wt % increases the lateral dimensions of the protrusions but not their density. dSP-B1-25 removes material from the fluid phase, as does SP-C, but the location and distribution of the material is quite different, leading to a smaller effect on the surface viscosity. Biophysical Journal 2005 89, 266-273DOI: (10.1529/biophysj.104.052092) Copyright © 2005 The Biophysical Society Terms and Conditions

Figure 8 Surface shear viscosity of Survanta (bovine lung extract supplemented with DPPC and PA) and Curosurf (porcine lung extract) replacement lung surfactants. Both show similar behavior to the lipid mixtures in that the surface viscosity strongly increases at a surface pressure (our magnetic needle surface viscometer can only measure up to ∼50 mN/m) corresponding to an SP-C induced percolation of the solid phase domains in the monolayer. Survanta has a lower transition surface pressure than Curosurf at 25°C, as the added PA and DPPC in Survanta increase the solid phase fraction as in Fig. 1 (inset). Raising the temperature of the Survanta film to 37° moves the transition to higher surface pressure, consistent with the observation that the solid phase fraction is smaller at higher temperatures (13,34). Raising the temperature of the Curosurf film to 37° presumably moves the transition to a surface pressure higher than could be measured using the magnetic needle viscometer (data not shown). Biophysical Journal 2005 89, 266-273DOI: (10.1529/biophysj.104.052092) Copyright © 2005 The Biophysical Society Terms and Conditions