Correlation of membrane protein conformational and functional dynamics

Membrane proteins (MPs) reside in the plasma membrane and perform various biological processes including ion transport, substrate transport, and signal transduction.*

Function-related conformational changes in membrane proteins occur in times scales ranging from nanoseconds to seconds.*

Obtaining time-resolved dynamic information of MPs in their membrane environment is still a major challenge.*

Although High Speed Atomic Force Microscopy (HS-AFM) images label-free samples such as DNA, soluble proteins, MPs, and intrinsically disordered proteins at ~1n~m lateral, ~0.1 nm vertical and ~100 ms temporal solution in aqueous environment and at ambient temperature and pressure, its temporal resolution is too slow to characterize many dynamic biological processes.*

In order to overcome this limitation Raghavendar Reddy Sanganna Gari, Joel José Montalvo-Acosta, George R. Heath, Yining Jiang, Xiaolong Gao, Crina M. Nimigean, Christophe Chipot and Simon Scheuring in their article Correlation of membrane protein conformational and functional dynamics use High Speed Atomic Force Microscopy Height Spectroscopy ( HS-AFM-HS) to characterize the microsecond timescale conformational changes of an integral-MP model system, i.e., the outer membrane protein G (OmpG) in a membrane environment.*

The positioning of the AFM tip is guided by HS-AFM imaging immediately before HS-AFM-HS-operation.*

NanoWorld Ultra-Short Cantilevers (USC) of the USC-F1.2-k0.15 type were used for the HS-AFM and HS-AFM-HS presented in the article.*

Figure 1 from “Correlation of membrane protein conformational and functional dynamics” by R.R. Sanganna Gari et al:
HS-AFM imaging of OmpG in lipid bilayers at pH 7.6 and pH 5.0.
a OmpG at pH 7.6 (Supplementary movie 1, left; frame rate: 200 ms per frame). A OmpG dimer is highlighted with dashed outline in all frames. Arrowheads in t = 11.6 s: Loop-6 fluctuating over the lumen. Arrowhead in t = 12.0 s: Fully open state. b Correlation average (n = 2752) of the HS-AFM movie frames (344 frames recorded over 68.8 s, full color scale: 0.0 nm < height < 1.25 nm, where the membrane level was set to 0.0 nm). c Correlation average of OmpG dimers. The topography outline (based on the molecular structure in 1e), serves as a visual guide to locate loop-6 and loop-2 in the topography and is highlighted by the dashed outline (the position of loop-6 is indicated by the asterisk based on its location in the structure (e)). Inner dashed outline show barrel lumen. d Standard deviation (std) map (n = 2752) from the averaging process in (b) (full color scale from blue to red: 0.05 nm < std < 0.19 nm) and topography outlines as in (c). e X-ray structure (PDB 2iwv) of the open OmpG conformation. Loop-6 (arrowhead L6) stands out of the image plane towards the viewer. Loop-2 (L2) forms a beta strand pointing away from the β-barrel, well detected by HS-AFM in the open state (b). f OmpG at pH 5.0 (Supplementary movie 1, right; frame rate: 200 ms per frame). A OmpG dimer is highlighted with dashed outline in all frames. g Correlation average (n = 2472) of the HS-AFM movie frames (309 frames recorded over 61.8 s, full color scale: 0.0 nm < height < 0.7 nm, where the membrane level was set to 0.0 nm). h Correlation average of OmpG dimers. For comparison, the topography outline of the open state (e) is shown (the position of loop-6 is indicated by the asterisk). i Standard deviation (std) map (n = 2472) from the averaging process in (g) (full color scale from blue to red: 0.04 nm < std < 0.07 nm) and topography outlines as in (h). j X-ray structure (PDB 2iww) of the closed OmpG conformation shown in the same orientation as in (e). Loop-6 (L6) folds over the β-barrel lumen in a lid-like manner. Loop-2 (L2) does not form a β-strand in the closed state, in agreement with absence of topography in this region in (h). Black dashed line: outline based on (e) for comparison.
NanoWorld USC-F1.2-k0.15 AFM probes were used for the HS-AFM.
Figure 1 from “Correlation of membrane protein conformational and functional dynamics” by R.R. Sanganna Gari et al:
HS-AFM imaging of OmpG in lipid bilayers at pH 7.6 and pH 5.0.
a OmpG at pH 7.6 (Supplementary movie 1, left; frame rate: 200 ms per frame). A OmpG dimer is highlighted with dashed outline in all frames. Arrowheads in t = 11.6 s: Loop-6 fluctuating over the lumen. Arrowhead in t = 12.0 s: Fully open state. b Correlation average (n = 2752) of the HS-AFM movie frames (344 frames recorded over 68.8 s, full color scale: 0.0 nm < height < 1.25 nm, where the membrane level was set to 0.0 nm). c Correlation average of OmpG dimers. The topography outline (based on the molecular structure in 1e), serves as a visual guide to locate loop-6 and loop-2 in the topography and is highlighted by the dashed outline (the position of loop-6 is indicated by the asterisk based on its location in the structure (e)). Inner dashed outline show barrel lumen. d Standard deviation (std) map (n = 2752) from the averaging process in (b) (full color scale from blue to red: 0.05 nm < std < 0.19 nm) and topography outlines as in (c). e X-ray structure (PDB 2iwv) of the open OmpG conformation. Loop-6 (arrowhead L6) stands out of the image plane towards the viewer. Loop-2 (L2) forms a beta strand pointing away from the β-barrel, well detected by HS-AFM in the open state (b). f OmpG at pH 5.0 (Supplementary movie 1, right; frame rate: 200 ms per frame). A OmpG dimer is highlighted with dashed outline in all frames. g Correlation average (n = 2472) of the HS-AFM movie frames (309 frames recorded over 61.8 s, full color scale: 0.0 nm < height < 0.7 nm, where the membrane level was set to 0.0 nm). h Correlation average of OmpG dimers. For comparison, the topography outline of the open state (e) is shown (the position of loop-6 is indicated by the asterisk). i Standard deviation (std) map (n = 2472) from the averaging process in (g) (full color scale from blue to red: 0.04 nm < std < 0.07 nm) and topography outlines as in (h). j X-ray structure (PDB 2iww) of the closed OmpG conformation shown in the same orientation as in (e). Loop-6 (L6) folds over the β-barrel lumen in a lid-like manner. Loop-2 (L2) does not form a β-strand in the closed state, in agreement with absence of topography in this region in (h). Black dashed line: outline based on (e) for comparison.
Figure 2 from “Correlation of membrane protein conformational and functional dynamics” by R.R. Sanganna Gari et al:
Single channel electrophysiology and HS-AFM height spectroscopy recordings of OmpG in lipid bilayers.
Representative 60-ms segments of OmpG single channel recordings at pH 7.6 (a) and pH 5.0 (b) at +40 mV membrane potential (longer traces in Supplementary Figs. 2 and 3). Cartoon representation of single channel recording experimental setup is shown in inset of (a). OmpG (yellow) in open state (PDB:2IWV) is placed in a lipid bilayer (green) surrounded by buffer (light blue shade) and potassium and chloride ions are shown as red and blue spheres. Red arrow indicates ion flow through OmpG in response to voltage application. Dwell time histograms of open and closed states at pH 7.6 (c) and pH 5.0 (d) from single-channel recordings (see Supplementary Table 1). Representative 60-ms segments of OmpG HS-AFM-HS recordings at pH 7.6 (e) and pH 5.0 (f) (longer traces in Supplementary Fig. 4). Cartoon representation of HS-AFM height spectroscopy experimental setup is shown in inset of (e). An oscillating AFM tip (orange) detects conformational changes of loop motion. Dwell time histograms of open and closed states at pH 7.6 (g) and pH 5.0 (h) from HS-AFM-HS recordings (Supplementary Table 2). In HS-AFM-HS the low state represents the open state, where the HS-AFM tip can descend into the β-barrel, and the high state represents the closed state, where loop-6 covers the beta barrel barring access of the HS-AFM tip to the cavity. All current-time and height-time traces were filtered at 20 kHz during analysis. The state dwell-time histograms are shown using log binning for better visualization of the components49. Red traces in (a) and (b) represent idealized current-time traces using clampfit software. Red traces in (e) and (f) represent idealized height-time traces using the STaSI algorithm (see Methods).
NanoWorld USC-F1.2-k0.15 AFM probes were used for the HS-AFM-HS.
Figure 2 from “Correlation of membrane protein conformational and functional dynamics” by R.R. Sanganna Gari et al:
Single channel electrophysiology and HS-AFM height spectroscopy recordings of OmpG in lipid bilayers.
Representative 60-ms segments of OmpG single channel recordings at pH 7.6 (a) and pH 5.0 (b) at +40 mV membrane potential (longer traces in Supplementary Figs. 2 and 3). Cartoon representation of single channel recording experimental setup is shown in inset of (a). OmpG (yellow) in open state (PDB:2IWV) is placed in a lipid bilayer (green) surrounded by buffer (light blue shade) and potassium and chloride ions are shown as red and blue spheres. Red arrow indicates ion flow through OmpG in response to voltage application. Dwell time histograms of open and closed states at pH 7.6 (c) and pH 5.0 (d) from single-channel recordings (see Supplementary Table 1). Representative 60-ms segments of OmpG HS-AFM-HS recordings at pH 7.6 (e) and pH 5.0 (f) (longer traces in Supplementary Fig. 4). Cartoon representation of HS-AFM height spectroscopy experimental setup is shown in inset of (e). An oscillating AFM tip (orange) detects conformational changes of loop motion. Dwell time histograms of open and closed states at pH 7.6 (g) and pH 5.0 (h) from HS-AFM-HS recordings (Supplementary Table 2). In HS-AFM-HS the low state represents the open state, where the HS-AFM tip can descend into the β-barrel, and the high state represents the closed state, where loop-6 covers the beta barrel barring access of the HS-AFM tip to the cavity. All current-time and height-time traces were filtered at 20 kHz during analysis. The state dwell-time histograms are shown using log binning for better visualization of the components49. Red traces in (a) and (b) represent idealized current-time traces using clampfit software. Red traces in (e) and (f) represent idealized height-time traces using the STaSI algorithm (see Methods).

*Raghavendar Reddy Sanganna Gari, Joel José Montalvo-Acosta, George R. Heath, Yining Jiang, Xiaolong Gao, Crina M. Nimigean, Christophe Chipot and Simon Scheuring
Correlation of membrane protein conformational and functional dynamics
Nature Communications volume 12, Article number: 4363 (2021)
DOI: https://doi.org/10.1038/s41467-021-24660-1

Please follow this external link to read the full article: https://rdcu.be/cAE2S

Open Access : The article “Correlation of membrane protein conformational and functional dynamics” by Raghavendar Reddy Sanganna Gari, Joel José Montalvo-Acosta, George R. Heath, Yining Jiang, Xiaolong Gao, Crina M. Nimigean, Christophe Chipot and Simon Scheuring is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this license, visit https://creativecommons.org/licenses/by/4.0/.

Millisecond dynamics of an unlabeled amino acid transporter

Excitatory amino acid transporters (EAATs) are important in many physiological processes and crucial for the removal of excitatory amino acids from the synaptic cleft.*

In the article “Millisecond dynamics of an unlabeled amino acid transporter “ Tina R. Matin, George R. Heath, Gerard H. M. Huysmans, Olga Boudker and Simon Scheuring develop and apply high-speed atomic force microscopy line-scanning (HS-AFM-LS) combined with automated state assignment and transition analysis for the determination of transport dynamics of unlabeled membrane-reconstituted GltPh, a prokaryotic EAAT homologue, with millisecond temporal resolution.*

Among the bulk and single-molecule techniques, high-speed atomic force microscopy ( HS-AFM ) stands out with its ability to provide real-time structural and dynamical information of single molecules. HS-AFM images label-free molecules under close-to-physiological conditions with ~0.1 nm vertical and ~1 nm lateral imaging resolution. Furthermore, HS-AFM has typically ~100 ms temporal resolution, giving access to structure–dynamics relationship of proteins, though the achievable imaging speed depends on sample characteristics like scan size and surface corrugation.

Recently in a quest to achieve higher temporal resolutions, the authors of the cited article used HS-AFM line scanning (HS-AFM-LS) for the analysis of single-protein dynamics. *

Line scanning, using a conventional AFM, has been used to study protein–protein interactions earlier. In HS-AFM-LS, the slow-scan axis (y-direction) is disabled. Therefore, instead of imaging an x/y-area, the scientists scan over one horizontal x-line several hundreds to thousands of times per second, thus reaching millisecond temporal resolution. The topographical readouts of this line are stacked one after another, resulting in kymographs of the dynamical behavior of the molecules. Therefore, HS-AFM-LS has between 2 and 3 orders of magnitude higher temporal resolution than HS-AFM imaging and should allow the detection of fast transporter dynamics and possible intermediate states that have so far escaped kinetic characterization. *

All AFM images presented in this study were taken using a HS-AFM operated in amplitude modulation mode (with typical free and setpoint amplitudes, Afree = 1.0 nm and Aset = 0.9 nm, respectively using optimized scan and feedback parameters. NanoWorld Ultra-Short Cantilevers ( NanoWorld’s AFM probe series especially dedicated for High Speed Scanning) of the USC-F1.2-k0.15 type were used. In the presented experiments, four different buffer conditions were used. *

As the authors state in their article they find that GltPh transporters can operate much faster than previously reported, with state dwell-times in the 50 ms range, and report the kinetics of an intermediate transport state with height between the outward- and inward-facing states. Transport domains stochastically probe transmembrane motion, and reversible unsuccessful excursions to the intermediate state occur. The presented approach and analysis methodology are generally applicable to study transporter kinetics at system-relevant temporal resolution.*

Figure 2 from “Millisecond dynamics of an unlabeled amino acid transporter” by Tina R. Matin et al.
HS-AFM line scanning (HS-AFM-LS): millisecond temporal resolution of unlabeled transporter dynamics.:
a HS-AFM image of a membrane packed with GltPh exposing the extracellular face before HS-AFM-LS (apo condition: 20 mM Tris-HCl, pH7.5, 150 mM KCl). Dashed lines indicate the position of the central scan line where subsequent HS-AFM-LS is performed. b Six seconds of a HS-AFM-LS kymograph with 3.3 ms line acquisition speed. Each transporter domain appears as a vertical line. c Projection (top) and height profile (bottom) of b. d HS-AFM image after HS-AFM-LS. The lateral position of recognizable features in a–d are indicated by arrowheads. e One second high-magnification views of dashed regions 1, 2, and 3 in b. Transport domain excursions to the inward-facing state appear as dark dwells along the vertical time axis. f Projection (top) and height profile (bottom) of e. Arrowheads indicate the position of the seven protomers in the kymograph (red: active protomer #5). g Height/time traces (gray) and state fits (red) of the active domain (protomer #5) in e. This figure is representative of the experimental sequence for the >50 replicates analyzed in this work.
NanoWorld Ultra-Short Cantilevers ( NanoWorld's AFM probe series especially dedicated for High Speed Scanning) of the USC-F1.2-k0.15 type (8 μm length, nominal spring constant of 0.15 N/m, nominal resonance frequency of ∼650 kHz and quality factor of ∼1.5 in buffer) were used.
Figure 2 from “Millisecond dynamics of an unlabeled amino acid transporter” by Tina R. Matin et al.
HS-AFM line scanning (HS-AFM-LS): millisecond temporal resolution of unlabeled transporter dynamics.:
a HS-AFM image of a membrane packed with GltPh exposing the extracellular face before HS-AFM-LS (apo condition: 20 mM Tris-HCl, pH7.5, 150 mM KCl). Dashed lines indicate the position of the central scan line where subsequent HS-AFM-LS is performed. b Six seconds of a HS-AFM-LS kymograph with 3.3 ms line acquisition speed. Each transporter domain appears as a vertical line. c Projection (top) and height profile (bottom) of b. d HS-AFM image after HS-AFM-LS. The lateral position of recognizable features in a–d are indicated by arrowheads. e One second high-magnification views of dashed regions 1, 2, and 3 in b. Transport domain excursions to the inward-facing state appear as dark dwells along the vertical time axis. f Projection (top) and height profile (bottom) of e. Arrowheads indicate the position of the seven protomers in the kymograph (red: active protomer #5). g Height/time traces (gray) and state fits (red) of the active domain (protomer #5) in e. This figure is representative of the experimental sequence for the >50 replicates analyzed in this work.

*Tina R. Matin, George R. Heath, Gerard H. M. Huysmans, Olga Boudker and Simon Scheuring
Millisecond dynamics of an unlabeled amino acid transporter
Nature Communications volume 11, Article number: 5016 (2020)
DOI: https://doi.org/10.1038/s41467-020-18811-z

Please follow this external link to read the full article: https://rdcu.be/cbuOU

Open Access : The article “Millisecond dynamics of an unlabeled amino acid transporter” by Tina R. Matin, George R. Heath, Gerard H. M. Huysmans, Olga Boudker and Simon Scheuring is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this license, visit https://creativecommons.org/licenses/by/4.0/.

The free energy landscape of retroviral integration

Retroviral integration, the process of covalently inserting viral DNA into the host genome, is a point of no return in the replication cycle. Yet, strand transfer is intrinsically iso-energetic and it is not clear how efficient integration can be achieved.*

In the article “The free energy landscape of retroviral integration” published in Nature Communications Willem Vanderlinden, Tine Brouns, Philipp U. Walker, Pauline J. Kolbeck, Lukas F. Milles, Wolfgang Ott, Philipp C. Nickels, Zeger Debyser and Jan Lipfert use biochemical assays, atomic force microscopy (AFM), and multiplexed single-molecule magnetic tweezers (MT) to study tetrameric prototype foamy virus (PFV) strand-transfer dynamics.*

Their finding that PFV intasomes employ auxiliary-binding sites for modulating the barriers to integration raises the question how the topology of higher-order intasomes governs integration of pathogenic retroviruses, most notably HIV. The single-molecule assays developed in this work are expected to be particularly useful to further unravel the complexity of this important class of molecular machines.*

The AFM images were recorded in amplitude modulation mode under ambient conditions and by using NanoWorld high resolution SuperSharpSiliconSSS-NCH cantilevers ( resonance frequency ≈300 kHz; typical end-radius 2 nm; half-cone angle <10 deg). Typical scans were recorded at 1–3 Hz line frequency, with optimized feedback parameters and at 512 × 512 pixels.*

Figure 2 e, f and g from “The free energy landscape of retroviral integration” by Willem Vanderlinden et al. 
(please refer to the full article for the complete figure 2  https://rdcu.be/b0R63 ) :
  e Atomic Force Microscopy image of intasomes incubated briefly (2 min) with supercoiled plasmid DNA, depicting a branched complex as found in ~50% of early complexes.
  f  Atomic Force Microscopy image of a bridging complex that dominates (~80%) the population of complexes at longer (>45 min) incubation. 
 g  Atomic Force Microscopy image of a gel-purified STC
Figure 2 e, f and g from “The free energy landscape of retroviral integration” by Willem Vanderlinden et al.
(please refer to the full article for the complete figure 2 https://rdcu.be/b0R63 ) :
 e AFM image of intasomes incubated briefly (2 min) with supercoiled plasmid DNA, depicting a branched complex as found in ~50% of early complexes.
 f AFM image of a bridging complex that dominates (~80%) the population of complexes at longer (>45 min) incubation.
g AFM image of a gel-purified STC

*Willem Vanderlinden, Tine Brouns, Philipp U. Walker, Pauline J. Kolbeck, Lukas F. Milles, Wolfgang Ott, Philipp C. Nickels, Zeger Debyser, Jan Lipfert
The free energy landscape of retroviral integration
Nature Communications volume 10, Article number: 4738 (2019)
DOI: https://doi.org/10.1038/s41467-019-12649-w

Please follow this external link to read the full article: https://rdcu.be/b0R63

Open Access The article “The free energy landscape of retroviral integration“ by Willem Vanderlinden, Tine Brouns, Philipp U. Walker, Pauline J. Kolbeck, Lukas F. Milles, Wolfgang Ott, Philipp C. Nickels, Zeger Debyser and Jan Lipfert is licensed under a Creative Commons Attribution 4.0 International License, which permits use, sharing, adaptation, distribution and reproduction in any medium or format, as long as you give appropriate credit to the original author(s) and the source, provide a link to the Creative Commons license, and indicate if changes were made. The images or other third party material in this article are included in the article’s Creative Commons license, unless indicated otherwise in a credit line to the material. If material is not included in the article’s Creative Commons license and your intended use is not permitted by statutory regulation or exceeds the permitted use, you will need to obtain permission directly from the copyright holder. To view a copy of this license, visit http://creativecommons.org/licenses/by/4.0/.