http://informahealthcare.com/mbc ISSN: 0968-7688 (print), 1464-5203 (electronic) Mol Membr Biol, 2014; 31(5): 141–151 ! 2014 Informa UK Ltd. DOI: 10.3109/09687688.2014.937469

REVIEW ARTICLE

Microscopy approaches to investigate protein dynamics and lipid organization Joanna M. Kwiatek, Elizabeth Hinde, and Katharina Gaus

Mol Membr Biol Downloaded from informahealthcare.com by University of Oklahoma on 08/17/14 For personal use only.

Centre for Vascular Research, Australian Centre for NanoMedicine and ARC Centre of Excellence in Advanced Molecular Imaging, University of New South Wales, Sydney, Australia Abstract

Keywords

The structure of cell membranes has been intensively investigated and many models and concepts have been proposed for the lateral organization of the plasma membrane. While proteomics and lipidomics have identified many if not all membrane components, how lipids and proteins interactions are coordinated in a specific cell function remains poorly understood. It is generally accepted that the organization of the plasma membrane is likely to play a critical role in the regulation of cell function such as receptor signalling by governing molecular interactions and dynamics. In this review we present different plasma membrane models and discuss microscopy approaches used for investigating protein behaviour, distribution and lipid organization.

Membrane order, microscopy techniques, protein interaction History Received 10 April 2014 Revised 9 June 2014 Accepted 17 June 2014 Published online 21 July 2014

Introduction

Models of the plasma membrane

Over the past decades, the lipid raft hypothesis has changed the way cell biologists view lipids and has postulated a role for lipid domains in protein organization and dynamics. These membrane domains are therefore often viewed as biophysically and biochemically discrete platforms. However, lipid domains remain controversial, mainly because biochemical assays are ambiguous and direct visualization has been proven difficult, suggesting that domains sizes are below the diffraction limit of optical microscopes. Therefore, single molecule and super resolution approaches are promising methods to investigate plasma membrane organization and they have the enormous potential to advance our knowledge of the lateral distribution of lipids and proteins in membranes. Here, we first summarize the models and concepts proposed for the organization of the plasma membrane, and then discuss the experimental evidence of heterogeneities in membrane order, the fluorescence techniques used to measure protein interactions and dynamics and finally outline the potential of super-resolution for membrane biology. In addition, we give an overview of the microscopy techniques discussed in Table 1.

In 1972 Singer and Nicolson proposed the fluid mosaic model (Figure 1A) that describes the plasma membrane as a twodimensional liquid in which proteins are homogenously distributed in the lipid bilayers. This model allows for rapid lateral diffusion of lipids and proteins in the membrane but essentially describes the membrane as a passive matrix (Singer & Nicolson, 1972). Cell membranes are however heterogeneous and have distinct and diverse membrane features and domains such as membrane protrusions, cellcell junctions or coated pits (Simons & Ikonen, 1997). In addition, molecular crowding, the existence of lipid and protein complexes that themselves may not be randomly distributed and local membrane deformation due to differences in lipid fluidity and proteins rigidity give the cell membrane a rather more complex architecture than simple lipid bilayers (Engelman, 2005). Simons and colleagues modified the fluid mosaic model with the concept of lipid-lipid interactions that led to the lipid raft hypothesis (Figure 1B) to describe protein trafficking in epithelial cells in 1997 (Simons & Ikonen, 1997). Seven years later, Kusumi et al. proposed the so-called picket-fence model (Figure 1C) in which transmembrane proteins (‘‘pickets’’) and the membrane skeleton (‘‘fences’’) confined the diffusion of membrane components due to collision (Kusumi et al., 2004). Membrane skeleton-dependent hop diffusion was directly observed for various receptors such as the transferrin receptor, a2-macroglobulin receptor, G-proteincoupled receptor or m-opioid receptor as test molecules (Akihiro & Sakot, 1996; Sako et al., 1998; Suzuki et al., 2005; Tomishige et al., 1998). When the lipid

Correspondence: Prof. Katharina Gaus, Centre for Vascular Research, Australian Centre for NanoMedicine and ARC Centre of Excellence in Advanced Molecular Imaging, University of New South Wales, Sydney, Australia. E-mail: [email protected]

Fluorescent molecules are irreversibly photobleached by a high power laser in the region of interest of the cell and the recovery of fluorescence from surrounding non-bleached fluorescence molecules monitored. Fluctuations in fluorescence intensity caused by the diffusion of fluorescent molecules through a small focal volume are observed. The corresponding autocorrelation function can be fitted to quantify diffusion coefficients and molecule concentrations within the focal volume.

A time series of frames is acquired using a raster scan, where pixels in each image are measured at different times and therefore the spatial correlation of the whole image contains details about dynamic processes occurring on different time scales.

Position of sparsely labelled single molecules is precisely mapped in time and space

Sample is scanned with two overlapping lasers. Molecules located in the outer part of the diffraction-limited excitation area are depleted by stimulated emission (doughnut shape beam) so that only

Fluorescent recovery after photobleaching (FRAP)

Raster image correlation spectroscopy (RICS)

Single particle tracking (SPT)

Stimulation emission depletion (STED)

Fluorescent correlation spectroscopy (FCS)

Fluorescent resonance energy transfer (FRET)

FLIM detects changes in the lifetime of a fluorescent dye or protein in each pixel of an image and the phasor approach transforms it into a fit free 2D coordinate system. A donor fluorophore non-radiatively transfers energy to an acceptor fluorophore by means of intermolecular long-range dipole-dipole coupling.

Principle

Phasor analysis to fluorescent lifetime microscopy (FLIM)

Method

Table 1. Summary of microscopy techniques.

Super-resolution technique with 30–150 nm resolution (depends on STED laser power)  Built on laser-scanning microscopy  Can be combined with FCS





 Labelling of individual molecules is required  Small number of molecules is measured  Requires complex analysis  Limited choice of fluorescent probes  Photo-bleaching due to high laser paper  Specialized microscope set-up

 Averaging over pixels, lines and frames within one measurement limits spatial resolution

 Molecular mobility and detection of transient confinement  Hop diffusion (requires ultrafast cameras)  Imaging with spatial super-resolution

 Measures protein behaviour such as formation of protein complexes and binding dynamics to an immobile or mobile structure.

 Measures:  Protein–protein interactions  Diffusion mode and coefficient  Chemical kinetics and photo-physical properties

(Geumann et al., 2010) (Halemani et al., 2010) (Singh et al., 2012) (Wurm et al., 2011)

(Kusumi et al., 2011)

(Digman et al., 2005) (Digman et al., 2009)

(Benda et al., 2003) (Hinde et al., 2013) (Clegg et al., 2005) (Enderlein & Gregor, 2005) (Bohmer, 2002) (Benda et al., 2005) (Benes et al., 2004)

J. M. Kwiatek et al.



 







 





 Does not provide spatial information  Long acquisition times  Data fitting and interpretation is complex

(Ishikawa-Ankerhold et al., 2012)

Single molecule technique (nM fluorophores concentrations) Small observation volume (510 ml) High temporal resolution with large range (msec-sec) Live cell measurements Can be applied to a single point measurement or line scan Relative simple method for diffusion measurements Can differentiate slow diffusion from binding events Large temporal range (msec-sec) Live cell measurements Applicable to confocal microscopy High spatial and temporal resolution Live cell measurements

 Ensemble measurement  Cannot reveal different diffusion modes

 Monitoring the mobility of proteins and fluorescent lipid analogues



   



(Clegg et al., 2005) (Varma & Mayor, 1998) (Kenworthy & Edidin, 1998) (Kenworthy et al., 2000) (Siegel, 2000)

 Measures protein-protein interactions  Protein association with domains  Receptor oligomerization

   

Can detect interactions 510 nm Suitable for live cell measurements Applicable to confocal microscopy Can be combined with FLIM Easy to perform Live cell measurements Applicable to confocal microscopy

  Photobeaching Probe orientation Dipole moment Must incorporate the distance of spacers

(Digman et al., 2008) (Owen et al., 2012b) (Golfetto et al., 2013)

 Membrane order measurements to detect lipid phases  Useful for FRET measurements

 Slow acquisition speed  Does not provided information on domain size and geometry

Does not require a priori knowledge of the number of components that need to be unmixed  Applicable to live cell imaging



Reference

Applications

Limitation

Advantage

Mol Membr Biol Downloaded from informahealthcare.com by University of Oklahoma on 08/17/14 For personal use only.

142 Mol Membr Biol, 2014; 31(5): 141–151

Imaging membrane dynamics

(Eggeling et al., 2009) (Mueller et al., 2011)

(Williamson et al., 2011) (Rossy et al., 2013) (Sengupta et al., 2011)

 Diffusion of fluorescent lipid analogues  First detection of differences in lipid diffusion and lipid nanoclusters

 Molecular quantification of protein distribution and clustering Stochastic optical reconstruction microscopy (STORM) and photoactivated localization microscopy (PALM)

Localization microscopy relies on the stochastic activation of conventional fluorophores (STORM) or photoswitchable fluorescent proteins (PALM). During one cycle, fluorescent molecules are photoswitched/activated, imaged and then photobleached, and the cycle repeated several thousand times.



Localization of individual molecules with 10–25 nm precision in lateral direction and 60–100 nm in axial direction  Single molecule data

 Limited choice of fluorescent probes  Photo-bleaching due to high laser paper  Specialized microscope set-up  Long acquisition times  Data fitting and interpretation is complex  Long acquisition time  Not suitable for all fluorophores  Photo-physical properties of fluorophores need to be understood for molecular counting  Complex analysis since each molecules needs to be fitted High spatial resolution of 30– 150 nm (due to reduced excitation volume)  High temporal resolution (5msec)  Live cell measurements

 STED-FCS

fluorophores from the central region of the excitation spot are able to emit fluorescence. Combines spatial resolution of STED with temporal resolution of FCS.

Mol Membr Biol Downloaded from informahealthcare.com by University of Oklahoma on 08/17/14 For personal use only.

DOI: 10.3109/09687688.2014.937469

143

L-a-dioleoylphosphatidylethanolamine (DOPE) was tagged with a colloidal gold particle, it localized in the outer leaflet of the plasma membrane and underwent hop diffusion that was dependent on the actin-based membrane skeleton (Fujiwara et al., 2002). It has also been shown that microdomain formation and sub-diffusion within the membrane might be enhanced by the association of actin filaments with the membrane bilayers (Ehrig et al., 2011; Garg et al., 2009). While the proposed membrane models suggested spatial organizations for proteins, lipids the membrane skeleton, the size of the lipid microdomains remains unknown. The Keystone Symposium in 2006 added a spatial dimension: ‘‘Small (10–200 nm), heterogeneous, highly dynamic, sterol- and sphingolipid-enriched domains that compartmentalize cellular processes. Small rafts can sometimes be stabilized to form larger platforms through proteinprotein and protein-lipid interactions’’ (Pike, 2006). The postulated small size and dynamic nature of rafts makes them challenging to observe directly, given that their dimensions are below the diffraction limit of light microscopy. The model called the lipid shell model (Figure 1D) suggests that 25% of the cell membrane surface contains proteins (20 000–35 000 proteins/mm2), which are surrounded by lipid rings (Jacobson et al., 2007). The number of integral membrane proteins in the endoplasmic reticulum and Golgi apparatus has been estimated as 30 000 molecules per mm2 (Quinn et al., 1984), the number of rhodopsin in the rod outer segment was assessed as 30 000 per mm2 (Liang et al., 2003; Liebman & Entine, 1974), and as a final example, the number of GPI-anchored folate receptor in caveolae following the addition of crosslinking antibodies was calculated as 32 000 per mm2 (Rothberg et al., 1990). Thus these estimations show that the plasma membrane contains a much higher protein concentration than artificial membranes. In 2011, Kusumi and colleagues also suggested that the organization of the plasma membrane occurs at different length scales. They distinguished three main compartments (a) ‘‘membrane compartments’’ which are 40–300 nm in diameter, (b) ‘‘raft domains’’ with a size between 2 and 20 nm in diameter, and the smallest (c) ‘‘dimers/oligomers and greater complexes of membraneassociated and internal membrane proteins’’ whose size is 3–10 nm (Kusumi et al., 2011). Similarly to co-existing membrane compartments, it should be pointed out that the above-presented models do not exclude each other (Figure 1E) and different domains may coexist in the plasma membrane.

Membrane order and the challenge to image lipid organization A central aspect of the lipid raft hypothesis is that the cell membrane exhibits two co-existing lipid phases: The liquid ordered phase (Lo) that mainly consists of saturated lipids and sterols, and liquid disordered phase (Ld) that is mainly made up by unsaturated lipids. The difficulty in directly imaging lipids is that a bulky fluorophore conjugated to a lipid often affects the partitioning and diffusion of the lipid analogue. Sezgin et al. showed that model systems such as giant unilameral vesicles (GUVs) are characterized by limited complexity and have different physical properties compared

Mol Membr Biol Downloaded from informahealthcare.com by University of Oklahoma on 08/17/14 For personal use only.

144

J. M. Kwiatek et al.

Mol Membr Biol, 2014; 31(5): 141–151

Figure 1. Different plasma membrane models. (A) Fluid mosaic model presents the plasma membrane as a two-dimensional liquid in which proteins (orange and blue shapes) are homogenously distributed in the lipid bilayers (dark blue). This model allows the possibility of rapid lateral diffusion of lipids and proteins within the membrane. (B) The lipid raft model proposes the existence of small regions (green shapes) enriched in cholesterol (red shapes) and sphingolipids. (C) The picket fence model suggests that transmembrane proteins act as ‘‘pickets’’ (orange pink and blue shapes) and the membrane skeleton as ‘‘fences’’ (dark orange net). The membrane skeleton confines the diffusion of membrane components due to collision. The membrane is depicted from cytosolic side. (D) The lipid shell model suggests that 25% of the cell membrane surface contains proteins (orange shapes) which are surrounded by lipid rings (purple and pink). (E) The models are exclusive and it is likely that different various membrane domains and compartments coexist in the plasma membrane.

to intact the plasma membrane or giant plasma membrane vesicles (GPMVs). They observed that fluorescent ‘‘raft’’ lipid analogues that did not partition into the Lo phase in GUVs did do so in GPMVs and concluded that the fluorophore geometry and chemical properties have a significant influence on the partitioning behaviour of lipid analogues (Sezgin et al., 2012). An alternative approach is to use fluorophores that partition into both phases and have fluorescent properties that are sensitive to membrane order. The fluorescence properties of the environmentally-sensitive membrane dyes are, for example, dependent on the polarity of their local solvent (Parasassi et al., 1998). Since penetration of polar water molecules into the non-polar bilayer interior depends on the degree of lipid packing, polaritysensitive dyes can sense the local molecular environment and thus help to distinguish Lo and Ld phases in the membrane (Figure 2A). One of the most common fluorophores for imaging lipid domains is Laurdan (6-lauryl-2-dimethylamino-naphthalene).

It is a polarity-sensitive dye created by Weber and Farris in 1979 as a derivative of Prodan (Weber & Farris, 1979). ˚ from the centre Laurdan’s fluorescent group is located 10 A of the bilayer (Antollini & Barrantes, 1998) while the Laurdan dipole molecules are aligned with the phospholipid head groups and the first carbon atom of the acyl chains in the phospholipid bilayers (Bagatolli et al., 1999). Laurdan’s fluorescent emission properties are dependent on the extent of water penetration into the bilayer (Figure 2B), which is different between ordered and disordered membrane domains (Gaus et al., 2003; Owen et al., 2010; Parasassi & Gratton, 1995; Parasassi et al., 1997, 1998). Laurdan’s lifetime is also membrane order-dependent (Figure 2C) where the Ld phase is characterized by a 4 ns lifetime and the Lo phase by a much longer lifetime of 6 ns (Parasassi et al., 1993). Another fluorescent probe often used to probe membrane order is the styryl di-4-ANEPPDHQ dye, which was designed as a voltage-sensitive dye to monitor electrical activity in cells and tissues (Obaid et al., 2004) but can also be used to differentiate lipid phases in model membranes and intact

Mol Membr Biol Downloaded from informahealthcare.com by University of Oklahoma on 08/17/14 For personal use only.

DOI: 10.3109/09687688.2014.937469

Figure 2. Spectral and lifetime methods of analysis to detect lipid phases. (A) The plasma membrane is depicted with the membrane dye Laurdan (orange shapes) in a random distribution between liquid disordered phase (Ld, blue lipids) and liquid ordered phase (Lo, green lipids). (B) Schematic emission spectra of lipid-phase sensitive dye, Laurdan, with green and purple spectra corresponding to Lo and Ld phases, respectively. (C) Typical fluorescence lifetime decays correspond to liquid ordered (dotted line) and liquid disordered (solid) phases, respectively. (D) Transformation of a spectrum or decay into a phasor (vector) plot. Spots represent data for Lo (lower) and Ld (upper) phases.

cells (Jin et al., 2005, 2006; Owen et al., 2012a,b). However di-4-ANEPPDHQ differs from Laurdan in the manner of membrane attachment and fluorescent lifetime. Di-4-ANEPPDHQ is attached to the membrane with its acyl heads being inserted deep into the membrane and its two positive charges confer a lesser ability to move between the inner and outer leaflets than Laurdan (Hof, 2012). Additionally, the lifetime of di-4-ANEPPDHQ (1.8 ns) is shorter in the Ld phase, and longer in the Lo phase (3.5 ns) compared to Laurdan (Owen et al., 2006). As an alternative to the previously described solvent-sensitive dyes is a novel group of voltage-sensitive hemicyanine dyes which exhibit electrochromism. These dyes can also be used to quantify membrane order and in contrast to di-4-ANEPPDHQ and Laurdan, hemicyanine dyes have a wide one-photon excitation spectrum that ranges between 440–670 nm (Kwiatek et al., 2013; Yan et al., 2012). The most common method to investigate membrane order is to quantify the shift in emission spectra of a membrane order sensitive dye (as depicted in Figure 2B for Laurdan) by Generalized Polarization (GP) analysis (Gaus et al., 2003; Parasassi et al., 1997, 1998). The Generalized Polarization function proposed by Parasassi et al. in 1990 is defined as the normalized fluorescence intensity ratio of the spectral channels that correspond to ordered and disordered phases. GP values range between 1 for very fluid membranes to +1 for the most ordered membrane and are a measure of the relative water content in the membrane, which is directly related to membrane order (Parasassi et al., 1997). GP values do not depend on the probe concentration (Parasassi et al., 1990). The main advantage of GP analysis is that the method is

Imaging membrane dynamics

145

Figure 3. Fluorescent approaches for detection of protein interactions and dynamics. (A) Fluorescence resonance energy transfer (FRET) with the most commonly used fluorescent protein pair, CFP-YFP. Middle scheme: FRET does not occur because of a large distance between the CFP and YFP proteins. Right scheme: FRET occurs when the distance (R0) between donor and acceptor is below 10 nm. (B) Fluorescence correlation spectroscopy (FCS). Middle scheme: The cross correlation function between two proteins is constant because the proteins are not assembled into complexes or the distance between them is too large. Right scheme: the cross correlation function indicates the presence of a protein complex. (C) Fluorescence recovery after photobleaching (FRAP). Middle scheme: a region of interest is photobleached region (black square). Right scheme: recovery of fluorescence intensity.

robust and data acquisition can be acquired on confocal and 2-photon microscopes. The phasor approach to fluorescence lifetime imaging microscopy (FLIM) is an alternative method to the GP analysis of emission spectra for monitoring the physical environment of membrane order sensitive dyes. FLIM detects changes in the lifetime of a dye in each pixel of an image (as depicted in Figure 2C for Laurdan) and the phasor approach transforms FLIM data into a two-dimensional coordinate system (Figure 2D), which allows us to distinguish between a mixture of immiscible phases and a homogeneous environment with intermediate order. It does not require a priori knowledge of the number of components that need to be unmixed (Digman et al., 2008). The major limitation of FLIM detection is the acquisition time. To determine the lifetime of a fluorophore with accuracy requires many more photons than are required for spectral imaging. This disadvantage makes FLIM detection of dynamic processes more difficult than spectral-based methods of detection (Owen et al., 2012b). Owen et al. used the FLIM phasor approach to unmix the lifetime from ordered and disordered phases and, by a linear combination of liquid ordered and disordered lifetimes, quantified the membrane coverage in live cell membranes stained with membrane order-sensitive dyes (Laurdan or di-4-ANEPPDHQ). From the lifetime changes that were

Mol Membr Biol Downloaded from informahealthcare.com by University of Oklahoma on 08/17/14 For personal use only.

146

J. M. Kwiatek et al.

acquired in one spectral channel, it was found that the majority of the plasma membrane is covered by the Lo phase (Owen et al., 2012b). More recently, Laurdan lifetime changes were measured with a modified two-channel detection scheme with the objective to spectrally separate detection of the Lo phase with channel 1 (460/80 nm) and the Ld phase with channel 2 (540/50 nm). Using the phasor analysis to analyze the lifetime recorded in each channel, this detection scheme resolved two different trajectories that were attributed to changes in polarity (channel 1) versus changes in cholesterol content (channel 2). This spectrally resolved approach to FLIM detection of Laurdan demonstrates an increasing dipolar relaxation effect cause by rotating water molecules in the membrane and the polarity of the environment changes the Laurdan lifetime. Based on these observations, Golfetto et al. claimed that Laurdan, when combined with spectrally resolved detection and the phasor FLIM approach can be used as a biosensor for cell membranes to detect and distinguish membrane fluidity from cholesterol content (Golfetto et al., 2013). The capability of the phasor method was further developed in recent work by Fereidouni et al. (2012). The phasor plot (Figure 2D) was extended to spectral imaging by translating the phasor plot from the lifetime domain to a wavelength representation. The spectral phasor method is more robust than standard GP analysis because the entire emission spectrum is considered through the collection of photons in many spectral channels rather than in just two channels, as is done with the traditional GP analysis. In summary, improvements have been made to quantify membrane order in cells more precisely and it is now possible to unmix the fluorescence signatures from lipid phase-sensitive dyes with the aid of phasor plots. However, these methods lack spatial resolution so that individual lipid domains and hence domain geometry and lifetimes cannot be revealed.

Protein interactions and dynamics: Energy transfer, fluctuation methods and photo-bleaching Given that most cell functions are controlled by protein behaviour (Rossy et al., 2014), one may wish to map protein distributions, interactions and dynamics within the membrane and infer the existence of domains from protein behaviour rather than image the membrane domains themselves. Prior to the advent of super-resolution microscopy (which is required to measure protein distribution and clustering on the nanometre scale, see below), advanced microscopy methods have been developed to quantify protein mobility, diffusion and interactions in cell membranes. The first method of detection discussed here employs fluorescence resonance energy transfer (FRET) (Figure 3A) as the readout of protein interaction. FRET is a non-radiative energy transfer from an excited chromophore (the donor, D) to another chromophore (the acceptor, A) by means of intermolecular long-range dipole-dipole coupling. The efficiency of the energy transfer critically depends on the Fo¨rster distance between the donor and acceptor and their fluorescence properties. The FRET efficiency of a donor-acceptor interaction is relative to the donor’s quantum yield and lifetime, the donor and acceptor

Mol Membr Biol, 2014; 31(5): 141–151

orientation (kappa factor), the degree of donor and acceptor spectral overlap, and is inversely proportional to the sixth power of distance between donor and acceptor, which usually has to be between 1 and 10 nm (Figure 3A, right). When a FRET interaction occurs, the donor fluorescence emission and lifetime are quenched, and the acceptor lifetime lengthened. Any one of these changes – quenching of the donor lifetime, decrease in the rate of donor photobleaching or sensitized acceptor fluorescence (Lippincott-Schwartz et al., 2001) – can be used for FRET detection but detection of changes in the donor lifetime has emerged as the preferred property to monitor FRET. This is because FLIM-FRET has the advantage that non-interacting species (un-quenched donor and acceptor) can be easily distinguished from the quenched donor lifetime. This is not possible with fluorescent emission-based methods since the donor and acceptor fluorescence spectra are linear combinations with or without FRET interactions (Chen et al., 2003; Hof, 2012). The most commonly used donor-acceptor pair consists of Cyan Fluorescent Protein (CFP) and Yellow Fluorescent Protein (YFP), in which the emission spectrum of the donor (CFP) highly overlaps with that of the acceptor (YFP). FRET is routinely used to investigate protein-protein interactions and membrane domain organization. For example, FRET was used to investigate Fas and how this cell surface receptor transduces an apoptotic signal upon binding to its trimeric ligand FasL. A FRET study revealed that Fas exists as a pre-assembled trimer on the cell surface, in contrast to previous assumptions that Fas is a monomer which assembles into trimers after FasL binding (Siegel, 2000). FRET was also used to investigate protein association with membrane domains. If a protein is clustered in a pre-existing lipid domain, the FRET efficiency between individual proteins would be controlled by the domain size and occupancy but relatively independent of the protein expression level, at least at high expression levels. However, if the protein is randomly distributed in the plasma membrane, the distance between proteins will strongly depend on the protein expression level. One out of three studies observed FRET efficiencies consistent with raft-induced clustering of the glycosylphosphatidylinositol-anchored protein (GPI-AP), folate receptor (Varma & Mayor, 1998), while two others claimed that the organization of GPI-AP and glycosphingolipids (GSL) was random (Kenworthy & Edidin, 1998; Kenworthy et al., 2000). The major advantage of FRET is that it can detect protein interactions because energy transfer is typically limited to distance of less than 10 nm. However, FRET is also limited by several photo-physical factors such as photobleaching, probe orientation and dipole moment (Clegg et al., 2005) and data acquisition is relatively slow so that short-lived interactions may be missed. An alternative method to probe protein diffusion and protein-protein interactions in live cells is fluorescence correlation spectroscopy (FCS) (Figure 3B). FCS monitors fluctuations in fluorescence intensity caused by the movement of fluorescent molecules through a small focal volume (He & Marguet, 2011; Magde et al., 1972). The autocorrelation function of these fluctuations can be fitted to quantify diffusion coefficients and molecule concentrations within the focal volume. FCS also gives information on molecular

Mol Membr Biol Downloaded from informahealthcare.com by University of Oklahoma on 08/17/14 For personal use only.

DOI: 10.3109/09687688.2014.937469

mobility and photophysical and photochemical reactions, such as translational diffusion, rotational diffusion (Benda et al., 2003), chemical kinetics (Clegg et al., 2005; Hinde et al., 2013), photon anti-bunching (Bohmer, 2002; Enderlein & Gregor, 2005), molecular interactions (Benda et al., 2005) and conformational fluctuations (Benes et al., 2004). FCS has been employed as a tool to investigate the mobility of proteins. Recently Hinde et al. combined twochannel FCS along a line scan (to introduce a spatial component) with FRET detection to investigate how protein-protein interactions change intracellular diffusion. The authors found from this multiplexed approach that the mechanism by which Rac1 and RhoA are organized and distributed upon cell migration is highly dependent on time and space and demonstrated that cross fluctuation analysis can reveal where and when protein-protein interactions occur in agreement with FRET (Hinde et al., 2013). While FCS is a single molecule technique and thus does not require high fluorophore concentrations, this technique does not provide spatial information (unless applied to line-scanning) and is restricted to observing protein dynamics without considering how these proteins may interact with their environment (Digman et al., 2005). Further, fitting and interpreting autocorrelation functions is not always straightforward and in cellular data, it is difficult to distinguish slow diffusing from transient binding. Digman et al. developed a fluctuation-based method of analysis called raster image correlation spectroscopy (RICS). This technique is based on a raster scan where pixels in an image are measured at different times (microsecond separation along a line, millisecond separation between lines and second separation between frames) and therefore the spatial correlation of the whole image contains details about dynamic processes occurring on different time scales. The RICS approach can be used for investigating protein behaviour such as diffusion in cells including measuring binding dynamics, interactions with subcellular structures and quantifying local concentrations. The typical result of a RICS analysis is a map of the spatial distribution of protein complexes in the cell and their diffusion and/or binding properties (Digman et al., 2005). The RICS approach has been extended to a crosscorrelation analysis (ccRICS) analogously to cross FCS, where two different proteins are monitored in two separate detection channels to assess interaction. ccRICS can be used to measure the diffusion of protein complexes, the spatial and temporal distribution of these complexes, and to evaluate the fractions of interacting molecules. In a recent study, Digman et al. used ccRICS to show the differences between cytoplasmic diffusion and binding of adhesion molecules. Moreover, the authors created maps of molecular interactions and their dynamics in the cell (Digman et al., 2009). RICS distinguishes between diffusion and binding which is the main advantage compared to other fluctuation methods. In the above-described techniques photobleaching was considered as a limitation, however, it can also be utilized to measure protein mobility in the membrane by an approach called fluorescence recovery after photobleaching (FRAP) (Figure 3C). In this method, fluorescent molecules are irreversibly photobleached by a high power laser in the region of interest (ROI) of the cell and the recovery of

Imaging membrane dynamics

147

fluorescence from surrounding non-bleached fluorescence molecules monitored. Analysis of FRAP data provides bulk information about protein kinetics including protein mobile fraction, Mf, and diffusion constant, D, as well as transport rate of binding/dissociation rate from other proteins. Mobility describes the fraction of fluorescent proteins which diffuse into the bleached region during the duration of the experiment, and the diffusion constant is a measure of the rate of protein movement in the absence of flow or active transport (Axelrod et al., 1976; White & Stelzer, 1999). Yechiel and Edidin used FRAP to investigate the lateral diffusion coefficient and the mobile fraction of a fluorescent lipid analogue and membrane proteins reviling the proof of proteinrich lipid domains existence in the plasma membrane and showing mobility of lipids and proteins in the domains (Yechiel & Edidin, 1987). Unlike FCS, FRAP is an ensemble measurement and therefore cannot reveal different diffusion modes at the single molecule level. FRAP parameters are influenced by a range of membrane scenarios and environments (Feder et al., 1996; Lippincott-Schwartz et al., 2001) and FRAP applications include monitoring the diffusion of plasma membrane proteins and phospholipids (IshikawaAnkerhold et al., 2012). However, given the small size of membrane domains and protein complexes, ensemble measurements like FRAP are difficult to interpret.

Nano-scale organization of membrane proteins and super-resolution microscopy One of the major limitations of optical microscopy was until recently the diffraction limit that restricts the resolution to 250 nm in lateral direction and 600 nm in axial direction in confocal microscopy with 500 nm excitation. The diffraction limit means that it is impossible to detect small domains or individual proteins in the crowded milieu of the plasma membrane because neighbouring domains and proteins are closer than the resolution limit. Given that the diffraction limit is a law of physics, it is not an overstatement that super-resolution microscopy that can overcome this limit and achieve spatial resolutions of 20 nm is anticipated to revolutionize cell and membrane biology. However, spatial resolution is only one limiting factor in mapping small and highly dynamic domains such as lipid rafts. Temporal resolution can also be a limiting. Kusumi et al. considered the smallest raft to be a complex of at least three molecules including one cholesterol molecule. He suggested that time can be used to distinguish small rafts from their surrounding environment. The lipid complexes must be stable for at least 10 ns since this is the time needed for a lipid to relocate to a neighbour position (Kusumi et al., 2011). It has been shown that cholesterol tends to create clusters with a lifetime from 1–100 ns in unsaturated lipid domains (Pasenkiewicz Gierula et al., 1991; Subczynski et al., 1991). These insights were not achieved with super-resolution microscopy but with single particle tracking (SPT) where only a few molecules are labelled and tracked very precisely in time and space. Suzuki et al. (2012) applied single-molecule tracking to study the organization of glycosylphosphatidylinositol-

Mol Membr Biol Downloaded from informahealthcare.com by University of Oklahoma on 08/17/14 For personal use only.

148

J. M. Kwiatek et al.

Mol Membr Biol, 2014; 31(5): 141–151

Figure 4. Super-resolution methods. (A) Features in the plasma membrane such as lipid rafts are below (green PSF) the diffraction limit (black PSF). (B) Stimulated emission depletion (STED). Left panel: PSF of confocal beam. Right panel: PSF of STED beam where in a doughnut shaped STED beam is overlaid onto a confocal beam (green) which reduces the focal volume. (C) Combination of STED and FCS. The autocorrelation function is shifted towards shorter time when STED is applied to FCS, which allows a precise quantification of the diffusion coefficient. (D) Photo-activated localization microscopy (PALM) or stochastic optical reconstruction microscopy (STORM). Single molecules are photoactivated so that there are a limited number of fluorescent events in each frame. The PSF of each fluorescent event is analysed to obtain the localizations of each molecule with nanometre precision. All localization are combined to produce a super-resolution image.

anchored proteins (GPI-APs), probably the most studied lipid raft marker in the plasma membrane. The authors described the dependency between the GPI-APs expression level and complex formation. At low expression levels, GPI-AP homodimers are formed, both through protein-protein interactions and stabilization by cholesterol/rafts. At higher expression levels, homodimer rafts are assembled into homo- or hetero-oligomer rafts of GPI-APs. It was also shown in this paper that actin filaments were not involved in homodimer raft formation. Sharma et al. used FRET to estimate the percentage of GPI-AP species that are presented in clusters as well as cluster sizes. Changes in anisotropy revealed that 20–40% of GPI-anchored proteins form clusters. Homo-FRET was used to reveal that on average, clusters contained four GPI-AP proteins while the remaining GPI-APs were monomers. Additionally the data showed that homodimers or homo-oligomers were preferentially formed over heterometric GPI-APs clusters. Interestingly this study also indicated that GPI-AP clustering was independent of the surface density of proteins (Sharma et al., 2004). A few years later, Goswami et al. (2008) used anisotropy to describe GPIAPs distributions at steady state. GPI-APs were organized at the nanoscale (510 nm) and optically resolvable scales (450 nm). It was assumed that the large scale domains may present properties reminiscent of lipid rafts (Simons & Vaz, 2004). The authors also emphasized that GPI-APs nanoclusters are dependent on ATP and actin filaments (Goswami et al., 2008). Recently the same group suggested that GPI-APs proteins might form nanoclusters due to cortical actin asters (Gowrishankar et al., 2012). The coupling of actin asters to a GPI nanocluster requires a membrane protein that becomes actively or passively trapped in the aster. New super-resolution microscopy techniques such as stimulated emission depletion (STED) (Figure 4B) (Hell &

Wichmann, 1994), photo-activated localization microscopy (PALM) (Betzig et al., 2006) or stochastic optical reconstruction microscopy (STORM) can now resolve structures smaller than Abbe’s diffraction limit (250 nm) and thus new insights about membrane organization and protein behaviour may be revealed. In STED nanoscopy, the fluorescent dye molecules located in the outer part of the diffraction-limited excitation area are depleted by stimulated emission so that only fluorophores from the central region of the excitation spot are able to emit fluorescence spontaneously (Hell & Wichmann, 1994). Images of actin and tubulin strikingly reveal the difference in resolving power of confocal versus STED imaging (Bu¨ckers et al., 2011; Kasper et al., 2010; Willig et al., 2007). STED imaging can also detect protein aggregates. For example, STED images showed that the SNAP25 protein (which regulates membrane fusion) forms clusters independently of the functionality of the SNARE motif and binds to clustered syntaxin (Geumann et al., 2010; Halemani et al., 2010), and mitochondrial proteins Tom20, VDAC1 and COX2 have been shown to be distributed in nanoscale clusters (Singh et al., 2012; Wurm et al., 2011). Recently, STED was combined with FCS to assess the diffusion of lipid molecules in nanoscale membrane domains. Reducing the focal volume with the STED beam (Figure 4C) revealed the mode of diffusion of phosphoethanolamine, which differed from that of sphingomyelin that becomes temporarily trapped in domains of 520–50 nm for 510 ms (Eggeling et al., 2009). Later work confirmed that shortchained unsaturated lipids diffuse more rapidly than longchain saturated lipids, as predicted by the lipid raft hypothesis (Mueller et al., 2011). Since STED-FCS is an extremely sensitive approach; only with SPT with ultrafast cameras and STED-FCS is it possible to detect nanosized clusters of sphingolipids, cholesterol and proteins in cell membranes

Imaging membrane dynamics

Mol Membr Biol Downloaded from informahealthcare.com by University of Oklahoma on 08/17/14 For personal use only.

DOI: 10.3109/09687688.2014.937469

(Mueller et al., 2011) but STED-FCS is technically not easy to implement and the instruments are expensive. STED also used high laser power that could damage the cells under investigation and to date, only a limited number of fluorophores are available for STED imaging and STED-FCS. PALM and STORM employ photoactivatable or photoswitchable fluorescent proteins (e.g. PA-GFP and mEOS) and photoswitchable dyes or dye pairs (e.g. Cy3–Cy5) to achieve super-resolution images of single molecules at high densities (Figure 4D). The major concept of these approaches is similar: Localization microscopy relies on the stochastic activation of fluorophores. During one cycle, fluorescent molecules are photoswitched/activated, imaged and then photobleached, and the cycle repeated several thousand times. Each fluorescent event is localized by fitting the point spread function (PSF) to the intensity profile. Merging all single molecule positions gives the final super-resolution image (Betzig et al., 2006; Rust et al., 2006). Our group used PALM to quantify the clustering of T cell signalling proteins (Rossy et al., 2013; Williamson et al., 2011) where clustering was mainly dependent on protein-protein interactions or caused by association with vesicles. Sengupta et al. provided a more complex picture of the protein organization in the plasma membrane over a wide range of spatial scales using pair correlation analysis of PALM data. They measured cluster size, density and variability of nanoscale clusters of GPI-AP, transmembrane proteins (Lat and vesicular stomatitis viral glycoprotein [VSVG]), and an inner leaflet lipidanchored protein (Lyn) and found that clusters differed in size and number of proteins. The radius of an average GPIAP cluster was smaller than 60 nm and contained on average 2–3 proteins. In contrast Lyn-PA-GFP molecules were formed into larger clusters (460 nm) and included three or more proteins. An average of 30% of Lat-PA-GFP clusters had radii above 100 nm with 2–40 proteins. Finally the size of VSVG-PAGFP clusters were also above 60 nm in radius which concentrate on average of 2.8  0.4 detected proteins (Sengupta et al., 2011). The major advantage of PALM and STORM is its ability to localize molecules with nanometre precision. This makes PALM and STORM very promising tools to investigate protein clustering below the diffraction limit but so far no consensus has emerged as to how proteins clusters in the plasma membrane are formed. Dynamic information is more difficult to obtain with single molecule localization microscopy and probes to map lipid organization are not yet available for these super-resolution microscopy approaches.

Conclusion Newly developed microscopy methods with high spatiotemporal resolution will afford us a better understanding of the complexity of the plasma membrane. These methods can return information about the lipid phase organization (e.g. Lo or Ld phase), lipid and protein interactions (e.g. clustering) and dynamics (e.g. diffusion). Moreover it is possible to combine two methods such as STED-FCS to exploit the temporal sensitivity of FCS and the improved spatial resolution of STED. However, currently there are distinct probes for lipid organization and protein behaviour such

149

as lipid phase sensitive probes (e.g. Laurdan), probes for protein-protein interactions (e.g. CFP-YFP FRET pair) or fluorescent proteins for single molecule localization microscopy (e.g. PA-GFP) and not every probe is compatible with each microscopy technique. This makes it difficult to use for example, the microscope method with highest spatial resolution, single molecule localization microscopy, for the detection of Lo and Ld phases in cell membrane. It also makes it difficult to compare and integrate data from different studies. There are still controversies and some discrepancies in the literature about fundamental parameters of lipid domains and how lipids influence protein dynamics. For example, the existence of hop-diffusion was suggested by Kusumi et al. using ultra-fast SPT (Kusumi et al., 2004) but was not observed by the Hell group with STED-FCS (Mueller et al., 2011). Similarly, whether protein dynamics is mainly regulated by protein-protein interactions (Douglass & Vale, 2005; Yokosuka & Saito, 2009) or also influenced by lipid domains and the lipid environment (Gaus et al., 2005; Holowka, 1997; Sohn et al., 2008; Tavano et al., 2006) is not clear at the moment. Ultimately, in cellular membranes, lipids and proteins must be studied together and this requires a better integration between fluorescent probes and advanced imaging technology.

Declaration of interest The authors report no conflicts of interest. The authors alone are responsible for the content and writing of the paper. The work was supported by the Australian Research Council, National Health and Medical Research Council of Australia (Program Grant 1037320).

References Akihiro K, Sakot Y. 1996. Cell surface organization by the membrane skeleton. Curr Opin Cell Biol 8:566–574. Antollini SS, Barrantes FJ. 1998. Disclosure of discrete sites for phospholipid and sterols at the protein-lipid interface in native acetylcholine receptor-rich membrane. Biochemistry 37: 16653–16662. Axelrod D, Koppel DE, Schlessinger J, Elson E, Webb WW. 1976. Mobility measurement by analysis of fluorescence photobleaching recovery kinetics. Biophys J 16:1055–1069. Bagatolli LA, Parasassi T, Fidelio GD, Gratton E. 1999. A model for the interaction of 6-lauroyl-2-(N,N-dimethylamino)naphthalene with lipid environments: Implications for spectral properties. Photochem Photobiol 70:557–564. Benda A, Benes M, Marec V, Lhotsky A, Hermens W, Heyrovsky J. 2003. How to determine diffusion coefficients in planar phospholipid systems by confocal fluorescence correlation spectroscopy. Langmuir 9:4120–4126. Benda A, Hof M, Wahl M, Patting M, Erdmann R, Kapusta P. 2005. TCSPC upgrade of a confocal FCS microscope. Rev Sci Instrum 76: 033106–4. Benes M, Billy D, Speijer H, Hof M, Hermens WT. 2004. Phospholipid membranes on mica, silica and glass. Langmuir 20:10129–10137. Betzig E, Patterson GH, Sougrat R, Lindwasser OW, Olenych S, Bonifacino JS, et al. 2006. Imaging intracellular fluorescent proteins at nanometer resolution. Science 313:1642–1645. Bohmer M. 2002. Time-resolved fluorescence correlation spectroscopy. Chem Phys Lett 353:439–445. Bu¨ckers J, Wildanger D, Vicidomini G, Kastrup L, Hell SW. 2011. Simultaneous multi-lifetime multi-color STED imaging for colocalization analyses. Opt Express 19:3130–3143.

Mol Membr Biol Downloaded from informahealthcare.com by University of Oklahoma on 08/17/14 For personal use only.

150

J. M. Kwiatek et al.

Chen Y, Mills JD, Periasamy A. 2003. Protein localization in living cells and tissues using FRET and FLIM. Differentiation 71:528–541. Clegg RM, Breusegam S, Barry N, Holub O, Govindjee G, Redford G. 2005. How Can one choose the best method for measuring FRET in a microscope with my biological system? Microsc Microanal 11: 272–273. Digman MA, Caiolfa VR, Zamai M, Gratton E. 2008. The phasor approach to fluorescence lifetime imaging analysis. Biophys J 94: L14–L16. Digman MA, Sengupta P, Wiseman PW, Brown CM, Horwitz AR, Gratton E. 2005. Fluctuation correlation spectroscopy with a laserscanning microscope: Exploiting the hidden time structure. Biophys J 88:L33–L36. Digman MA, Wiseman PW, Horwitz AR, Gratton E. 2009. Detecting protein complexes in living cells from laser scanning confocal image sequences by the cross correlation raster image spectroscopy method. Biophys J 96:707–716. Douglass AD, Vale RD. 2005. Single-molecule microscopy reveals plasma membrane microdomains created by protein-protein networks that exclude or trap signaling molecules in T cells. Cell 121:937–950. Eggeling C, Ringemann C, Medda R, Schwarzmann G, Sandhoff K, Polyakova S, et al. 2009. Direct observation of the nanoscale dynamics of membrane lipids in a living cell. Nature 457:1159–1162. Ehrig J, Petrov EP, Schwille P. 2011. Near-critical fluctuations and cytoskeleton-assisted phase separation lead to subdiffusion in cell membranes. Biophys J 100:80–89. Enderlein J, Gregor I. 2005. Using fluorescence lifetime for discriminating detector afterpulsing in fluorescence-correlation spectroscopy. Rev Sci Instrum 76:033102–5. Engelman DM. 2005. Membranes are more mosaic than fluid. Nature 438:578–80. Feder TJ, Brust-Mascher I, Slattery JP, Baird B, Webb WW. 1996. Constrained diffusion or immobile fraction on cell surfaces: A new interpretation. Biophys J 70:2767–2773. Fereidouni F, Bader AN, Gerritsen HC. 2012. Spectral phasor analysis allows rapid and reliable unmixing of fluorescence microscopy spectral images. Opt Express 20:12729–12741. Fujiwara T, Ritchie K, Murakoshi H, Jacobson K, Kusumi A. 2002. Phospholipids undergo hop diffusion in compartmentalized cell membrane. J Cell Biol 157:1071–1081. Garg S, Tang JX, Ru¨he J, Naumann C. 2009. Actin-induced perturbation of PS lipid-cholesterol interaction: A possible mechanism of cytoskeleton-based regulation of membrane organization. J Struct Biol 168:11–20. Gaus K, Chklovskaia E, Fazekas de St Groth B, Jessup W, Harder T. 2005. Condensation of the plasma membrane at the site of T lymphocyte activation. J Cell Biol 171:121–131. Gaus K, Gratton E, Kable EPW, Jones AS, Gelissen I, Kritharides L, Jessup W. 2003. Visualizing lipid structure and raft domains in living cells with two-photon microscopy. Proc Natl Acad Sci USA 100: 15554–15559. Geumann U, Scha¨fer C, Riedel D, Jahn R, Rizzoli S. 2010. Synaptic membrane proteins form stable microdomains in early endosomes. Microsc Res Tech 73:606–617. Golfetto O, Hinde E, Gratton E. 2013. Laurdan fluorescence lifetime discriminates cholesterol content from changes in fluidity in living cell membranes. Biophys J 104:1238–1247. Goswami D, Gowrishankar K, Bilgrami S, Ghosh S, Raghupathy R, Chadda R, et al. 2008. Nanoclusters of GPI-anchored proteins are formed by cortical actin-driven activity. Cell 135:1085–1097. Gowrishankar K, Ghosh S, Saha SCR, Mayor S, Rao M. 2012. Active remodeling of cortical actin regulates spatiotemporal organization of cell surface molecules. Cell 149:1353–1367. Halemani ND, Bethani I, Rizzoli SO, Lang T. 2010. Structure and dynamics of a two-helix SNARE complex in live cells. Traffic 11: 394–404. He H-T, Marguet D. 2011. Detecting nanodomains in living cell membrane by fluorescence correlation spectroscopy. Ann Rev Phys Chem 62:417–436. Hell SW, Wichmann J. 1994. Breaking the diffraction resolution limit by stimulated emission: Stimulated-emission-depletion fluorescence microscopy. Opt Lett 19:780–782. Hinde E, Digman MA, Hahn KM, Gratton E. 2012. Millisecond spatiotemporal dynamics of FRET biosensors by the pair correlation

Mol Membr Biol, 2014; 31(5): 141–151

function and the phasor approach to FLIM. Proc Natl Acad Sci USA 110:135–40. Hof M. 2012. Springer Series on Fluorescence Methods and Applications. Berlin: Springer Heidelberg. Holowka D. 1997. Compartmentalized activation of the high affinity immunoglobulin E receptor within membrane domains. J Biol Chem 272:4276–4280. Ishikawa-Ankerhold HC, Ankerhold R, Drummen GPC. 2012. Advanced fluorescence microscopy techniques – FRAP, FLIP, FLAP, FRET and FLIM. Molecules 17:4047–4132. Jacobson K, Mouritsen O, Anderson R. 2007. Lipid rafts: At a crossroad between cell biology and physics. Nat Cell Biol 9:7–14. Jin L, Millard AC, Wuskell JP, Clark HA, Loew LM. 2005. Cholesterolenriched lipid domains can be visualized by di-4-ANEPPDHQ with linear and nonlinear optics. Biophys J 89:L04–6. Jin L, Millard AC, Wuskell JP, Dong X, Wu D, Clark HA, Loew LM. 2006. Characterization and application of a new optical probe for membrane lipid domains. Biophys J 90:2563–2575. Kasper R, Harke B, Forthmann C, Tinnefeld P, Hell SW, Sauer M. 2010. Single-molecule STED microscopy with photostable organic fluorophores. Small 6:1379–1384. Kenworthy AK, Petranova N, Edidin M. 2000. High-resolution FRET microscopy of cholera toxin B-subunit and GPI-anchored proteins in cell plasma membranes. Mol Biol Cell 11:1645–1655. Kenworthy A, Edidin M. 1998. A glycosylphosphatidylinositol-anchored protein at the apical surface of MDCK cells examined at a resolution ˚ using imaging fluorescence resonance energy. J Cell Biol of 5100 A 142:69–84. Kusumi A, Koyama-Honda I, Suzuki K. 2004. Molecular dynamics and interactions for creation of stimulation-induced stabilized rafts from small unstable steady-state rafts. Traffic 5:213–230. Kusumi A, Suzuki KGN, Kasai RS, Ritchie K, Fujiwara TK. 2011. Hierarchical mesoscale domain organization of the plasma membrane. Trends Biochem Sci 36:604–615. Kwiatek JM, Owen DM, Abu-Siniyeh A, Yan P, Loew LM, Gaus K. 2013. Characterization of a new series of fluorescent probes for imaging membrane order. PLoS One 8:e52960. Liang Y, Fotiadis D, Filipek S, Saperstein DA, Palczewski K, Engel A. 2003. Organization of the G protein-coupled receptors rhodopsin and opsin in native membranes. J Biol Chem 278:21655–21662. Liebman P, Entine G. 1974. Lateral diffusion of visual pigment in photoreceptor disk membranes. Science 185:457–459. Lippincott-Schwartz J, Snapp E, Kenworthy A. 2001. Studying protein dynamics in living cells. Nat Rev Mol Cell Biol 2:444–456. Magde D, Elson E, Webb W. 1972. Thermodynamic fluctuations in a reacting system – measurement by fluorescence correlation spectroscopy. Phys Rev Lett 29:705–708. Mueller V, Ringemann C, Honigmann A, Schwarzmann G, Medda R, Leutenegger M, et al. 2011. STED nanoscopy reveals molecular details of cholesterol- and cytoskeleton-modulated lipid interactions in living cells. Biophys J 101:1651–1660. Obaid AL, Loew LM, Wuskell JP, Salzberg BM. 2004. Novel naphthylstyryl-pyridium potentiometric dyes offer advantages for neural network analysis. J Neurosci Meth 134:179–190. Owen D, Lanigan PMP, Dunsby C, Munro I, Grant D, Neil MAA, et al. 2006. Fluorescence lifetime imaging provides enhanced contrast when imaging the phase-sensitive dye di-4-ANEPPDHQ in model membranes and live cells. Biophys J 90:L80–L82. Owen D, Magenau A, Majumdar A, Gaus K. 2010. Imaging membrane lipid order in whole, living vertebrate organisms. Biophys J 99: L7–L9. Owen D, Rentero C, Magenau A, Abu-Siniyeh A, Gaus K. 2012a. Quantitative imaging of membrane lipid order in cells and organisms. Nat Protoc 7:24–35. Owen D, Williamson D, Magenau A, Gaus K. 2012b. Sub-resolution lipid domains exist in the plasma membrane and regulate protein diffusion and distribution. Nat Commun 3:1256. Parasassi T, De Stasio G, d’Ubaldo A, Gratton E. 1990. Phase fluctuation in phospholipid membranes revealed by Laurdan fluorescence. Biophys J 57:1179–1186. Parasassi T, Gratton E. 1995. Membrane lipid domains and dynamics as detected by Laurdan fluorescence. J Fluoresc 5:59–69. Parasassi T, Gratton E, Yu WM, Wilson P, Levi M. 1997. Two-photon fluorescence microscopy of laurdan generalized polarization domains in model and natural membranes. Biophys J 72:2413–2429.

Mol Membr Biol Downloaded from informahealthcare.com by University of Oklahoma on 08/17/14 For personal use only.

DOI: 10.3109/09687688.2014.937469

Parasassi T, Krasnowska EK, Bagatolli L, Gratton E. 1998. Laurdan and prodan as polarity-sensitive fluorescent membrane probes. J Fluoresc 8:365–373. Parasassi T, Ravagnan G, Rusch RM, Gratton E. 1993. Modulation and dynamics of phase properties in phospholipid mixtures detected by Laurdan fluorescence. Photochem Photobiol 57:403–410. Pasenkiewicz Gierula M, Subczynski WK, Kusumi A. 1991. Influence of phospholipid unsaturation on the cholesterol distribution in membranes. Biochimie 73:1311–1316. Pike LJ. 2006. Rafts defined: a report on the Keystone Symposium on Lipid Rafts and Cell Function. J Lipid Res 47:1597–1598. Quinn P, Griffiths G, Warren G. 1984. Density of newly synthesized plasma membrane proteins in intracellular membranes II. Biochemical studies. J Cell Biol 98:2142–2147. Rossy J, Ma Y, Gaus K. 2014. The organisation of the cell membrane: do proteins rule lipids? Curr Opin Chem Biol 20C:54–59. Rossy J, Owen DM, Williamson DJ, Yang Z, Gaus K. 2013. Conformational states of the kinase Lck regulate clustering in early T cell signaling. Nat Immunol 14:82–89. Rothberg KG, Ying YS, Kamen B a, Anderson RG. 1990. Cholesterol controls the clustering of the glycophospholipid-anchored membrane receptor for 5-methyltetrahydrofolate. J Cell Biol 111:2931–2938. Rust MJ, Bates M, Zhuang X. 2006. Imaging by stochastic optical reconstruction microscopy (STORM). Nature Methods 3:793–795. Sako Y, Nagafuchi A, Tsukita S, Takeichi M, Kusumi A. 1998. Cytoplasmic regulation of the movement of E-cadherin on the free cell surface as studied by optical tweezers and single particle tracking: corralling and tethering by the membrane skeleton. J Cell Biol 140: 1227–1240. Sengupta P, Jovanovic-Talisman T, Skoko D, Renz M, Veatch SL, Lippincott-Schwartz J. 2011. Probing protein heterogeneity in the plasma membrane using PALM and pair correlation analysis. Nat Methods 8:969–975. Sezgin E, Levental I, Grzybek M, Schwarzmann G, Mueller V, Honigmann A, et al. 2012. Partitioning, diffusion, and ligand binding of raft lipid analogs in model and cellular plasma membranes. Biochim Biophys Acta 1818:1777–1784. Sharma P, Varma R, Sarasij RC, Ira RC, Gousset K, Krishnamoorthy G, et al. 2004. Nanoscale organization of multiple GPI-anchored proteins in living cell membranes. Cell 116:577–589. Siegel RM. 2000. Fas preassociation required for apoptosis signaling and dominant inhibition by pathogenic mutations. Science 288: 2354–2357. Simons K, Ikonen E. 1997. Functional rafts in cell membranes. Nature 387:569–572. Simons K, Vaz WLC. 2004. Model systems, lipid rafts, and cell membranes. Ann Rev Biophys Biomol Struct 33:269–295. Singer S, Nicolson G. 1972. The fluid mosaic model of the structure of cells membranes. Science 175:720–731.

Imaging membrane dynamics

151

Singh H, Lu R, Rodrı´guez PFG, Wu Y, Bopassa JC, Stefani E, Toro L. 2012. Visualization and quantification of cardiac mitochondrial protein clusters with STED microscopy. Mitochondrion 12:230–236. Sohn HW, Tolar P, Pierce SK. 2008. Membrane heterogeneities in the formation of B cell receptor-Lyn kinase microclusters and the immune synapse. J Cell Biol 182:367–379. Subczynski WK, Hyde JS, Kusumi A. 1991. Effect of alkyl chain unsaturation and cholesterol intercalation on oxygen transport in membranes: a pulse ESR spin labeling study. Biochemistry 30: 8578–8590. Suzuki K, Ritchie K, Kajikawa E, Fujiwara T, Kusumi A. 2005. Rapid hop diffusion of a G-protein-coupled receptor in the plasma membrane as revealed by single-molecule techniques. Biophys J 88: 3659–3680. Suzuki KGN, Kasai RS, Hirosawa KM, Nemoto YL, Ishibashi M, Miwa Y, et al. 2012. Transient GPI-anchored protein homodimers are units for raft organization and function. Nat Chem Biol 8:774–783. Tavano R, Contento RL, Baranda SJ, Soligo M, Tuosto L, Manes S, Viola A. 2006. CD28 interaction with filamin-A controls lipid raft accumulation at the T-cell immunological synapse. Nat. Cell Biol 8: 1270–1276. Tomishige M, Sako Y, Kusumi A. 1998. Regulation mechanism of the lateral diffusion of band 3 in erythrocyte membranes by the membrane skeleton. J Cell Biol 142:989–1000. Varma R, Mayor S. 1998. GPI-anchored proteins are organized in submicron domains at the cell surface. Nature 394:798–801. Weber G, Farris FJ. 1979. Synthesis and spectral properties of a hydrophobic fluorescent probe: 6-propionyl-2-(dimethylamino)naphthalene. Biochemistry 18:3075–3078. White J, Stelzer E. 1999. Photobleaching GFP reveals protein dynamics inside live cells. Trends Cell Biol 9:61–65. Williamson DJ, Owen DM, Rossy J, Magenau A, Wehrmann M, Gooding JJ, Gaus K. 2011. Pre-existing clusters of the adaptor Lat do not participate in early T cell signaling events. Nat Immunol 12: 655–662. Willig KI, Harke B, Medda R, Hell SW. 2007. STED microscopy with continuous wave beams. Nat Methods 4:915–918. Wurm CA, Neumann D, Lauterbach M, Harke B, Egner A, Hell SW, Jakobs S. 2011. Nanoscale distribution of mitochondrial import receptor Tom20 is adjusted to cellular conditions and exhibits an inner-cellular gradient. Proc Natl Acad Sci USA 108:13546–13551. Yan P, Acker CD, Zhou W-L, Lee P, Bollensdorff C, Negrean A, et al. 2012. Palette of fluorinated voltage-sensitive hemicyanine dyes. Proc Natl Acad Sci USA 109:20443–8.. Yechiel E, Edidin M. 1987. Micrometer-scale domains in fibroblast plasma membranes. J Cell Biol 105:755–760. Yokosuka T, Saito T. 2009. Dynamic regulation of T-cell costimulation through TCR-CD28 microclusters. Immunol Rev 229:27–40.

Microscopy approaches to investigate protein dynamics and lipid organization.

The structure of cell membranes has been intensively investigated and many models and concepts have been proposed for the lateral organization of the ...
2MB Sizes 2 Downloads 4 Views