Next Article in Journal
Blood Transcript Biomarkers Selected by Machine Learning Algorithm Classify Neurodegenerative Diseases including Alzheimer’s Disease
Next Article in Special Issue
Phosphatidic Acid Accumulates at Areas of Curvature in Tubulated Lipid Bilayers and Liposomes
Previous Article in Journal
Druggable Metabolic Vulnerabilities Are Exposed and Masked during Progression to Castration Resistant Prostate Cancer
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Review

Small-Angle Neutron Scattering for Studying Lipid Bilayer Membranes

Neutron Scattering Division, Oak Ridge National Laboratory, Oak Ridge, TN 37831, USA
Biomolecules 2022, 12(11), 1591; https://doi.org/10.3390/biom12111591
Submission received: 22 September 2022 / Revised: 18 October 2022 / Accepted: 26 October 2022 / Published: 29 October 2022
(This article belongs to the Special Issue Recent Developments in Biophysical Studies of Cell Membranes)

Abstract

:

Featured Application

This manuscript has been authored by UT-Battelle, LLC under Contract DE-AC05-00OR22725 with the U.S. Department of Energy (DOE). The U.S. government retains and the publisher, by accepting the article for publication, acknowledges that the US government retains a nonexclusive, paid-up, irrevocable, worldwide license to publish or reproduce the published form of this manuscript, or allow others to do so, for U.S. government purposes.

Abstract

Small-angle neutron scattering (SANS) is a powerful tool for studying biological membranes and model lipid bilayer membranes. The length scales probed by SANS, being from 1 nm to over 100 nm, are well-matched to the relevant length scales of the bilayer, particularly when it is in the form of a vesicle. However, it is the ability of SANS to differentiate between isotopes of hydrogen as well as the availability of deuterium labeled lipids that truly enable SANS to reveal details of membranes that are not accessible with the use of other techniques, such as small-angle X-ray scattering. In this work, an overview of the use of SANS for studying unilamellar lipid bilayer vesicles is presented. The technique is briefly presented, and the power of selective deuteration and contrast variation methods is discussed. Approaches to modeling SANS data from unilamellar lipid bilayer vesicles are presented. Finally, recent examples are discussed. While the emphasis is on studies of unilamellar vesicles, examples of the use of SANS to study intact cells are also presented.

Graphical Abstract

1. Introduction

Biological membranes are composed of a wide variety of lipids and proteins and are integral to cellular function. Eukaryotic membranes contain a high fraction of PC and PE lipids (a list of abbreviations is provided immediately prior to the References), along with other phospholipids, cholesterol, sphingomyelin, and other spingolipids [1]. In contrast, bacterial membranes tend to be richer in PG and PE lipids, although their composition in different species can vary considerably [2,3]. Example lipids are shown in Figure 1. The complex mixture of lipids and proteins assembles into a well-defined and selectively permeable boundary between the interior and exterior of the cell and between structures within the cell. The membrane structure is not uniform across its thickness. The inner and outer leaflets of the membrane differ in composition [4], which is actively maintained by the cell [5]. Additionally, the structure of the membrane is not uniform within the plane of the membrane [6]. The interaction between the constituents of the bilayer provides the membrane with structure and a well-defined function, and no component works in isolation. The complexity of the membrane as well as the large range of length and time scales involved in its structure and function present a challenge to researchers who wish to understand it.
Multiple biophysical characterization techniques are required to understand the structure and function of biomembranes, because no single experimental technique is able to provide all of the required information. For example, protein conformation in the membrane environment can be probed by circular dichroism spectroscopy [7,8], solid-state magic-angle-spinning NMR [9], Fourier transform infrared spectroscopy [10], and fluorescence spectroscopy [11]. Structures in the membrane can be visualized at micron length scales by microscopy [12,13] and can be inferred from fluorescence spectroscopy [14]. NMR spectroscopy can be used to determine lipid acyl chain order parameters [15], and it can be used to determine the lipid bilayer structure [16].
X-ray and neutron scattering techniques are extensively used to determine the biomembrane structure. While diffraction can provide atomic-resolution information, SANS and SAXS are powerful tools for investigating structures in materials that lack crystallographic order at length scales from 10 Å to well over 1000 Å. Reflectometry is also a powerful technique for the study of biomembrane structure. The most detailed information comes from using highly oriented samples to perform lamellar diffraction [17,18,19,20,21,22,23,24,25]. Reflectometry also uses planar samples, but the samples typically studied are a single lipid bilayer adsorbed onto a highly flat substrate [26,27,28,29,30,31]. Reflectometry probes length scales similar to those of SANS, and it can be used in conjunction with SANS [32]. Hydrated dispersions and multilamellar vesicles, which provide powder diffraction data, generally provide less detailed structural information [33,34,35]. However, this is not necessarily the case because the entire scattering curve, if it is of high enough quality, can be fit for high-quality structural information [36,37]. As a result, SAXS studies often employ multilamellar vesicles. SANS instruments are not well-suited to these studies due to their lower ability to resolve features in the signal [38].
SANS and SAXS are powerful tools for studying lipid bilayer membranes in solution when prepared as unilamellar vesicles. Unilamellar vesicles are attractive because they are structurally similar to cell membranes, they are conceptually simple, and they are always in a fully hydrated state. They are also relatively easy to prepare using well-established methods such as rehydration from dried lipid films [39,40], extrusion [41,42], sonication [43,44], and electroformation [45]. Literature describing the difficulties that can be encountered during the preparation of unilamellar vesicles also exists [46,47,48]. As a result, experiments utilizing SANS and SAXS to investigate lipid bilayer membranes are readily undertaken by researchers from different backgrounds.
Here, a review of SANS for investigating lipid bilayer membranes is presented, with a particular focus on studies of unilamellar lipid bilayer vesicles. In addition to providing a brief overview of small-angle scattering, contrast variation methods for studying membranes are presented. Approaches to SANS data analysis for these systems are presented as well, which may also be applied to SAXS data. Finally, recent examples of the use of SANS to study lipid bilayer membranes are provided, with some examples of SANS being used in studies of intact, living biological membranes that help to show how studies of model systems can expand into studies of natural systems.

2. Small-Angle Neutron Scattering

Consider neutrons with wavelength λ , which interact with a sample and scatter from it, as shown schematically in Figure 2. The neutrons are incident on the sample with momentum k i and exit the sample with momentum k f , which is an angle 2 θ from the incident neutron beam direction. Note that only elastic scattering is being considered, and thus k i = k f = 2 π / λ . SANS data are measured as a function of the momentum transfer q = k f k i , which has a magnitude of q = ( 4 π / λ ) s i n ( θ ) and is referred to as q. It is related to the length scale d that is probed through d = 2 π / q . By measuring at small angles, i.e., small q, structures at large length scales relative to the neutron wavelength can be studied.
When neutrons interact with the atoms in the sample, they radiate out as circular waves (recall the wave–particle duality) that are centered on each atom. Small-angle scattering results from the interference of these waves. The measured small-angle scattering intensity of a material, regardless of whether neutrons or X-rays are used, is given by Equation (1) [49,50].
I ( q ) = | V ρ ( r ) e i q · r d 3 r | 2
Here, ρ ( r ) is the density of the material property that interacts with the probe. For neutrons, the density of interest is the neutron scattering length density, referred to as the SLD, which is discussed in Section 3. The integral is taken over the volume of the sample exposed to the incident beam V.
When applied to dilute solutions, such as lipid bilayer vesicles in water, Equation (1) can be re-written as Equation (2) [49,50].
I ( q ) = n p | V p [ ρ ( r ) ρ s ] e i q · r d 3 r | 2
In Equation (2), ρ ( r ) is the SLD inside the particle in solution, while ρ s is the SLD of the solvent. The integral is now taken over the volume of the particle V p . The denotes averaging over all orientations of the particle with respect to the incident beam. Equation (2) can be alternatively expressed according to Equation (3).
I ( q ) = n p | F ( q ) | 2 S ( q )
F ( q ) is referred to as the form factor, and S ( q ) is the structure factor. A lot of F ( q ) s are analytical functions [51], which is very helpful when performing data analysis. If there are no interparticle interactions, S ( q ) = 1 .

3. Contrast Variation

SANS has advantages that make it excellently suited for the study of membranes. The low-energy neutrons used for SANS do not cause radiation damage to most samples, which cannot be said for the X-rays used in SAXS experiments. However, the unique sensitivity of neutrons to light elements and the isotopes of elements is their greatest advantage. Unlike X-rays, neutrons interact with most elements on the periodic table as well as with their isotopes, with a comparable strength that is represented by the coherent neutron scattering length b c o h , often simply referred to as the neutron scattering length. The b c o h for the elements and several of their isotopes were tabulated by Sears [52], and those of interest for studies of membranes are presented in Table 1. Note that not only does 1 H have a significantly different neutron scattering length as compared to 2 H, but the sign of b c o h also differs. 1 H is one of the few elements that has a negative b c o h . The large difference between the b c o h of 1 H and 2 H makes contrast variation methods for biological materials relatively easy [53].
Contrast variation leverages the difference in scattering length density of different compounds in a structure, such as lipids, nucleic acids, and proteins, as well as the ability to label these materials with 2 H to highlight specific ones. The existence of contrast between materials in the sample means there are materials with different SLDs in the sample, and this makes it possible to differentiate portions of structures within the intact whole. The difference in SLD between two materials is often simply referred to as the contrast. Then, one varies the relative contrast between the parts of a structure by using different mixtures of H 2 O and 2 H 2 O as the solvent.
To calculate the SLD, ρ , of a molecule, one needs to know its atomic and isotopic composition and its molecular volume V. Then, the SLD can be calculated using Equation (4). The summation is over all i atoms in the molecule.
ρ = i b i , c o h V
It is necessary to account for hydrogen exchange with certain chemical groups, such as hydroxyl groups, when samples are in mixtures of H 2 O and 2 H 2 O. The molecular volume can be calculated from the molecular weight and density, if these are known. The volumes of various lipids as well as many head groups have been presented in a broad body of literature [19,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83]. It is important to note that not all of the published values for a given lipid, HG, or acyl chain agree. Temperature-dependent lipid volumes have also been measured [74,76,82,83,84]. Several example lipid HGs, acyl chains, and other example compounds encountered in studies of biomembranes are presented in Table 2. The convention for separating the head group from the non-polar groups of the lipids is presented in Figure 3 of the review by Marsh [80]. The SLDs of proteins, DNA, and RNA can be readily calculated from the sequence and the molecular volume. Information about the amino acid and nucleic acid volumes, exchangeable protons, and scattering lengths in H 2 O and 2 H 2 O are provided in Table 2 of the excellent review by Jacrot [85] and in the more recent comprehensive review by Breyton and coworkers [86]. The online activation calculator provided by the National Institute of Standards and Technology’s Center for Neutron Research can also provide the SLDs of proteins, RNA, and DNA [87]. The peptide melittin is also presented in Table 2 as an example.
In lipid bilayer membranes containing proteins or more than a single species of lipid, contrast variation methods make it possible to observe the protein in the vesicle or to see how the various lipid species are distributed. Lipids, proteins, and nucleic acids have intrinsic contrast without deuterium labeling [53]. To obtain a sufficient contrast between different lipids, deuterium-labeled lipids are needed. Fortunately, a wide variety of deuterium-labeled lipids, including those that are fully deuterated and those with specific chemical groups deuterated, are available for purchase from vendors. However, typically, only saturated acyl chains can be acquired in deuterated form from vendors. When studying lipid bilayer vesicles composed of a single lipid, one need not resort to contrast variation methods. SAXS is sufficient because there is electron density contrast between the lipid HGs and the acyl chains. SAXS data can also serve as an additional contrast point in a contrast variation study.
An example of contrast variation and selective deuterium labeling can be seen in Figure 3. Figure 3A presents SANS data from vesicles made of 2 H 54 -DMPC (i.e., the acyl chains are fully deuterated) at a 7:3 molar ratio and DMPE. The SLDs of 2 H 54 -DMPC and DMPE are 5.394 × 10 6 Å 2 and 0.323 ×   10 6 Å 2 , respectively, while the average SLD of the acyl chains of this composition is 4.673 ×   10 6 Å 2 . The samples were measured in three different mixtures of H 2 O and 2 H 2 O: 50%, 70%, and 90% 2 H 2 O, which have SLDs of 2.912 × 10 6 Å 2 , 4.301 × 10 6 Å 2 , and 5.690 × 10 6 Å 2 , respectively. The 50% 2 H 2 O SANS data result from the interplay of competing positive and negative contrasts. The 70% 2 H 2 O SANS data are dominated by the scattering from the HGs because the average SLD of the chains are nearly the same as that of the solvent. The 90% 2 H 2 O SANS data are dominated by the distribution of the DMPE in the vesicles.
In Figure 3B, SANS and SAXS data collected from vesicles made of a 7:3 molar mixture of DMPC and DMPG in 2 H 2 O are presented. The SANS data are from a previous study [89]. The hydrogenated lipids in 2 H 2 O have excellent contrast for SANS, but the details are lost because the contrast between the solvent and both the acyl chains and the HGs have the same sign and are relatively large, resulting in a less feature-rich data set. The contrast in the SAXS data set comes from the electron density of water being intermediate to those of the electron-rich HGs and acyl chains. The much better q-resolution (i.e., the uncertainty in q arising from the instrument) provided by SAXS instruments as compared to SANS instruments can also be seen in the sharper features in the SAXS data relative to those of the 50% 2 H 2 O SANS data in Figure 3A, even though the interplay of the contrasts in the samples are very similar.

4. Data Analysis

SANS data analysis and modeling for unilamellar lipid bilayer vesicles are reasonably straightforward. A variety of approaches that provide different levels of detail are used. A modified form of the Guinier analysis can be used to determine the bilayer thickness. Simple core–shell models are frequently used, but improved approaches that provide even more detailed structural information are available. Molecular dynamics simulations can also play a role. Here, a brief overview of data analysis methods for the interpretation of SANS data from unilamellar lipid bilayer vesicles is presented.
It is possible to determine the thickness of the bilayer by fitting the SANS data with the use of the Kratky–Porod approximation [90,91,92,93,94,95,96], which is similar to the Guinier analysis [50]. If the thickness of the vesicle, T v , is much smaller than the radius of the vesicle, R v , then the data, I ( q ) , can be expressed as I K P using the approximation shown in Equation (5).
I K P ( q ) I ( q = 0 ) q 2 e q 2 R T 2
R T is the radius of gyration of the thickness, and T v = R T 12 . R T is determined by fitting a straight line to L n ( q 2 I ( q ) ) vs. q 2 . The approximation is only valid for a region of q that is bound by structural features of the vesicles. More specifically, data fitting using the approximation shown in Equation (5) must be restricted to 1 / R v q 1 / R T [95]. Alternatively, the linear fit can be restricted to 2 π / S v q 1 / R T , where S v = 4 π R v 2 is the surface area of the vesicle [90,91,92]. An example of this kind of analysis is shown in Figure 4. When contrast variation methods are applied, additional details about the structure of the bilayer can be extracted from this analysis [90,91,92,93,94,95,96].
The most straightforward approach for modeling the structure of the vesicle is through the use of a core–shell model to fit the data [51]. The model intensity profile is shown in Equation (6) [51].
I ( q ) = A i = 1 N ( ρ i + 1 ρ i ) 4 π R i 3 3 3 ( sin q R i q R i cos q R i ) ( q R i ) 3 2 + B
In the equation, ρ 1 is the SLD of the core, which is assumed to be the same as the solvent ( ρ N + 1 ), while the ρ i represents the SLDs of the various shells in the model. The radii of the various shells, starting from the radius of the core, are represented by R i . If the thickness of a shell in the model is T i , then R i + 1 = R i + T i . The radius of the core of the vesicle, which is R 1 , may be polydisperse, but the radii of the layers are normally assumed to have a single value. There is always one more term in the summation than there are shells in the model. A and B are a multiplicative constant and an additive baseline, respectively. A is related to the sum of the products of the SLDs and volumes of the shells in the model, and the concentration of the vesicles. It is often easier to simply allow it to be a free parameter. B accounts for the incoherent scattering from hydrogen, but it also accounts for background subtraction artifacts during data reduction.
The number of shells to be used depends on the complexity of the bilayer structure. If three shells are used, then they represent an inner HG region, the acyl chain region, and the outer HG region. When contrast variation methods are used in samples containing mixtures of lipids, such as for the samples shown in Figure 3A, four shells are reasonable because the hydrogenated and deuterated lipids may not be equally distributed between the leaflets of the bilayer. The four shells represent the inner HG, the inner acyl chain region, the outer acyl chain region, and the outer HG, shown schematically in Figure 5.
A variety of approaches exist for determining the T i and ρ i of the various shells in the model. The values of T i and ρ i are allowed to be free parameters, but they may be constrained by the known physical parameters and the composition of the samples. Modeling SANS data from lipid bilayer vesicles in this way is relatively simple. No new software must be developed for the data analysis because a suitable model is provided in the SasView software package [98]. However, relating the resulting SLDs of the layers to their composition is not trivial.
An extension of this modeling approach is carried out by calculating the SLDs of the layers in the model from the known composition of the samples [99,100,101,102]. Consider a vesicle made from an undeuterated lipid that has a volume of V L 1 and HG and acyl chain volumes of V H G 1 and V A C 1 , respectively, and a deuterium-labeled lipid with a volume of V L 2 and HG and acyl chain volumes of V H G 2 and V A C 2 , respectively. Assume that the compositions of the inner and outer leaflets of the lipid bilayer are identical; hence, using a three-shell model is appropriate. Let the SLDs of the HGs and acyl chains of the lipids be ρ H G 1 , ρ A C 1 , ρ H G 2 , and ρ A C 2 . Denote the molar fraction of the unlabeled lipid with f, and let the number of waters per lipid associated with the inner and outer HG regions of the bilayer be N w , i n and N w , o u t , respectively. The SLD of water is denoted by ρ w , and the volume of a water molecule is V w . The SLDs of the three layers in the model are calculated using Equations (7)–(9).
ρ H G , i n = f ρ H G 1 V H G 1 + ( 1 f ) ρ H G 2 V H G 2 + ρ w N w , i n V w f V H G 1 + ( 1 f ) V H G 2 + N w , i n V w .
ρ A C = f ρ A C 1 V A C 1 + ( 1 f ) ρ A C 2 V A C 2 f V A C 1 + ( 1 f ) V A C 2 .
ρ H G , o u t = f ρ H G 1 V H G 1 + ( 1 f ) ρ H G 2 V H G 2 + ρ w N w , o u t V w f V H G 1 + ( 1 f ) V H G 2 + N w , o u t V w .
Note that it is not necessary to include a factor of 2 in Equation (8) because both the numerator and denominator would contain the factor of 2.
In light of the wealth of information that is available regarding lipid volumes and the volumes of acyl chains and HGs [19,54,55,56,57,58,59,60,61,62,63,64,65,66,67,68,69,70,71,72,73,74,75,76,77,78,79,80,81,82,83], it is possible to provide additional constraints on the structure of the lipid bilayer that is represented by the shells; this is termed the self-consistent slab model [103,104,105]. Recently, a version of the model was made available by Tan and coworkers [106] for use in the Sasview data analysis software package [98]. The self-consistent slab model shares features of other core–shell models that use the composition to determine the SLDs of the various layers in the model, but the model also uses known volumes for components of the bilayer to define the layer thicknesses and to provide a physically realistic structure. Here, one way of defining a self-consistent model is described using the schematic in Figure 6.
Assume that the bilayer has a single component. The area per lipid, A L , and the volume of the acyl chains, V H C , are used to determine the thickness of the acyl chain region, T H C , according to Equation (10). The model described here does not take the curvature of the vesicle into account, as in other self-consistent models [103,104,105,106], because it assumes that A L is constant along the entire lipid molecule.
T H C = V H C A L
The thickness of the HG region of the bilayer, T H G , may be a free parameter if it is constrained such that A L T H G cannot be smaller than the volume of the lipid HG, V H G , or that it can be fixed [82,107,108], which is performed here. Alternatively, the model can be parameterized using the number of waters associated with the HG, N w , which is then used to determine T H G [106]. Then, the thicknesses of the layers are given in Equation (11).
T 1 = T H G ; T 2 = 2 T H C ; T 3 = T H G
A L also determines the number of waters associated with the HGs in the model here by using Equation (12).
N w = A L T H G V H G V w
The SLDs of the layers are calculated from the sample composition using Equations (7)–(9).
As the desire for more detailed structural information about the structure of lipid bilayers has increased, more detailed structural models have been developed. A notable result was the scattering density profile or SDP model [108,109]. Analytical functions are used to represent the position distributions of various chemical groups in the bilayer structure [109,110]. SDP models use constraints on the volumes of the groups that the lipid structure is divided into to ensure a physically realistic model. Fogarty and coworkers also expanded the modeling approach to an atomistic level of detail [111]. An online SDP modeling tool that allows anyone to fit their data using the approach was recently made available [112].
MD simulations can also be used for relating small-angle scattering data from lipid bilayer membranes to the membrane structure. Programs such as nMoldyn [113] or Sassena [114], which calculate X-ray and neutron scattering intensities directly from an MD trajectory, will not be discussed. Instead, methods more suitable for determining the transbilayer SLD profile are discussed here. Early work used simulations to calculate diffraction form factors, subsequently using the data to refine simulation parameters [115]. While the interpretation of diffraction data continued to be the main focus [110,116,117], the techniques later expanded to include SANS and SAXS [82,83,103,108,109,118,119,120,121]. MD simulations are less commonly used than the modeling methods described above because they are computationally intensive and require an MD expert. The position distributions of the various chemical groups in the bilayer can be determined directly from the simulation. Then, these distributions are used to calculate the bilayer SLD profile, which is subsequently used to calculate the SANS intensity profile that is compared with the measured data. Simulating an entire vesicle at an atomic resolution is prohibitive, but it can be accomplished using coarse-grained simulation methods such as the MARTINI approach [122]. Carillo and coworkers also developed a coarse-grained simulation method for modeling neutron scattering data [123].
Analyzing or modeling SANS data from vesicles that are not uniform in the plane of the bilayer—such as when domains form in a system that has been made using multiple lipids and when selective deuteration is employed—presents different challenges. System-specific modeling approaches are often required. When no obvious lateral domains are visible in the data, but there is evidence of two thicknesses in the same vesicle, then data can be modeled using an additive superposition of two bilayer models [100,124,125]. The Kratky–Porod approach to fitting data can reveal the presence of domains when contrast variation methods are utilized [96]. If a picture of the structures formed is desired, more complex models for the vesicle that account for the domains and the correlations between them are required. For example, Heberle and coworkers used a Monte Carlo approach to simulate the SANS data from vesicles that possessed large domains [126]. The models accurately reproduced the impact of the large domains on the SANS data and revealed how the domains depended on the composition of the vesicles and the temperature.
It is important to note that one of the challenges that faces modeling for SAS data analysis is the over-fitting of data due to limited information content [127,128]. It is possible to have more than one model fit the data equally well, particularly when too many free parameters are used. For example, a four-shell model, in which the layer thicknesses and SLDs are allowed to be free parameters, has ten free parameters (see Equation (6)) or more if the core radius and layers are polydisperse and the widths of the distributions are not fixed. It is generally preferable to use as few free parameters as possible. Self-consistent slab models and the SDP model help to reduce the number of free parameters required during fitting. Using molecular dynamics simulations to determine the SLD profile across the membrane also constrains the material spatial distributions. Laterally inhomogeneous systems may require several free parameters during fitting, and care must be taken to ensure that the results are robust.

5. Examples

Many researchers have employed SANS to study the structure of lipid bilayer vesicles. It has been used as a tool for understanding membrane structure, primarily in model systems containing a limited mix of lipid species, although more complex membranes have also been investigated. The technique has also found application in understanding the interaction of lipid bilayers with peptides and proteins, small molecules, and polymers. Intact cells and organisms have also been studied. Here, a survey of more recent examples taken from the large body of literature that employed SANS to study lipid bilayer membranes is presented.

5.1. SANS for Lipid Bilayer Structure Determination

SANS is a valuable tool for structural studies, both in concert with diffraction and by itself. Several different single-component PC lipid bilayers have been characterized by SANS, including the saturated DLPC [119,129], DMPC [91,95,119,129,130,131], DPPC [119,132,133,134,135], DSPC [119,129,136], and DPhPC [74,81]. PC lipids that have monounsaturated [74,119,137] and polyunsaturated [138] acyl chains have been studied as unilamellar vesicles using SANS. Various PG lipids [82], PE lipids [83], and PS lipids [120] have also been characterized. Cardiolipin bilayer membranes have been studied [139] as well as those formed by sphingomyelins [140], the GM3 ganglioside [141], and lipopolysaccharides extracted from cells [142,143]. More novel lipids have been studied using SANS, such as fully artificial ones [144] and siloxane phosphocholines [145]. SANS also provides the important membrane thickness needed for understanding the dynamics of lipid bilayers that have been measured by NSE [146,147,148,149,150,151,152,153,154], although the information reported about the structure may not be as detailed as what is found in a study dedicated to membrane structure.
More complex lipid mixtures composed of multiple phospholipids, or that incorporate cholesterol or sphingomyelin, have also been studied with the use of SANS. Cholesterol is an important constituent of mammalian cell membranes, and it is also medically important. The impact of cholesterol on the structure of phospholipid bilayers was the subject of several studies [84,155,156,157,158,159,160,161]. SANS has been used to determine the solubility limit of cholesterol in POPC and POPS unilamellar vesicles and found it to be quite different [162]. SANS and MD have been used to understand membranes made of ether lipids and cholesterol [163] as well as membranes from the stratum corneum [164]. The impact of ergosterol on DPPC membranes was determined using SANS [165]. DMPC/DMPG/DHPC bilayer vesicles that spontaneously form from a bicelle state have also been characterized [166,167,168]. A mixture of PC lipids with sphingomyelin was studied, which resulted in bilayers showing the significant interdigitation of the inner and outer leaflets of the membrane [169]. Mimics of bacterial membranes containing PE, PG, and various cardiolipins were probed by SANS [170]. DMPC-ceramide mixtures produced smaller vesicles than when only DMPC was present [171]. It is also possible to investigate vesicles in which the inner leaflet of the membrane does not have the same composition as the outer one [172,173,174,175,176].
SANS with contrast variation methods are excellent tools for investigating the formation of laterally separated domain structures in membranes such as rafts [6]. SANS has revealed lateral inhomogeneities in vesicles containing gangliosides [96,177], sphingomyelin [178], vesicles made of lipid mixtures that have very different acyl chain lengths [126], and simple phospholipid–cholesterol mixtures [179]. Domains have also been observed using SANS using contrast variation methods in lipid vesicles made of POPC and DPPC [180]. Ceramides were found to impact the stability of domains formed in mixtures containing DPPC, DOPC, and cholesterol [181]. Ahmadi and coworkers used SANS and QENS to characterize mixtures of DPPC, DOPC, and cholesterol and of DPPC, POPC, and cholesterol [182]. Both mixtures became phase-separated at a temperature below the gel phase transition of DPPC and displayed raft-like domains, shown schematically with the SANS data in Figure 7. Similarly, contrast-matching methods can be used to visualize domains in vesicles, as was carried out for a mixture of POPC and DSPC [183].

5.2. SANS for Studying Lipid Bilayer Interactions between Peptides and Proteins

SANS with contrast variation is very useful for investigating the interaction of peptides and proteins with lipid bilayer membranes. The impact of the peptide on the structure of the bilayer can be determined, as well as how the peptide impacts the distribution of lipids if selective deterium labeling is employed. Several groups have used SANS in studies of peptides with lipid bilayers. Amyloid plaque-forming peptides such as the Alzheimer’s amyloid β peptide [184], α -synuclein [185], and the islet amyloid polypeptide [186] were investigated in lipid bilayer vesicles using SANS. The interaction and self-assembly of the peptides were studied. SANS and contrast variation methods are also excellent for studying the interaction of membrane-active and antimicrobial peptides with lipid bilayers because additional details about the interaction can be revealed when selective deuteration is employed [99,100,124,187,188,189]. It was possible to determine where indolicidin associates with vesicles [189] and how aurein 1.2 interacts with DMPC/DMPG vesicles [188]. The distribution of cholesterol was studied by Qian and coworkers when alamethicin [100] and melittin [124] interacted with vesicles made of DMPC and cholesterol, which indicated that both peptides disrupted how cholesterol was distributed in the vesicles. The movement of charged lipids between the inner and outer leaflets of the vesicles were also observed for alamethicin and melittin [99,187]. In Figure 8, the extent of charged lipid redistribution caused by melittin in DMPC/DMPS vesicles is shown schematically and demonstrates the sensitivity of contrast variation methods [187]. SANS can also be used to study the interaction of full-length proteins with lipid bilayer vesicles. Examples that employed contrast-matching techniques to minimize the scattering from the lipid include a study of cytochrome c oxidase [190] and the Bax/tBid system [191]. However, contrast matching is not required, as was demonstrated by Doktorova and coworkers in their study of the HIV-1 Gag protein [192].
SANS is often used as a complementary tool in neutron spectroscopy studies of the impact of peptides on the dynamics of lipid bilayer membranes. QENS provides information about local dynamics in membranes, and it has been used with complementary SANS experiments to study the amyloid β peptide fragment (1–40) in DMPG membranes [187], which can be seen in Figure 9, as well as alamethicin and melittin in DMPC membranes [194]. Fusion peptides from the SARS-CoV-2 virus were found by QENS to make membranes more rigid [195]. At longer length- and time-scales, NSE and SANS are also valuable tools for understanding how peptides impact lipid bilayer membranes. In addition to revealing the impact of the SARS-CoV-2 fusion peptides on membranes [195], the impact of the HIV-1 fusion peptide has been studied [89], as has the amyloid β peptide [196]. Kelley and coworkers also determined how the antimicrobial peptides alamethicin and gramicidin slow down the collective dynamics of membranes, indicating that the membranes become stiffer in response to the peptides [197].

5.3. SANS for Lipid Bilayer Interactions with Other Materials

Several groups have investigated the interaction of lipid bilayers with compounds that are not proteins or peptides. The interaction of a variety of small molecules with membranes has been probed by SANS, including sugar [198,199,200], various alcohols produced during biofuel production [201], DMSO [92,202,203,204], glycerol [205], and carboranes [206]. Detergents [91,207,208,209], alkanes [210,211,212,213], and fatty acids [214] have also been studied. Anesthetics that interact directly with membranes were investigated in lipid bilayer membranes using SANS [215,216,217]. SANS improved the understanding of how vitamin E [218] and melatonin [219,220], both important for human health, interact with membranes. Compounds from food, some of which have medicinal effects, interacting with membranes have been studied using SANS, such as aescin from horse chestnut [221,222,223,224], glycyrrhizin from licorice [225], and epigallocatechin-3-gallate from green tea [226]. Furthermore, medicines have been studied. Painkillers, including ibuprofen [222,227] and acetominophen [228], interacting with lipid bilayers were probed with SANS. SANS was also used to study the interaction of the antifungal amphotericin B [229] and the antimicrobial HT61 [230] with lipid bilayer vesicles. Khadka and coworkers used SANS to characterize how the anticancer drug tamoxifen associates with lipid bilayers [231].
Polymers are of interest for drug delivery, in which the cell membrane plays an important role. They are also ubiquitous in daily life, and their interaction with membranes is of interest in a general sense. Poly(ethylene glycol)-modified lipids and their impact on vesicle structure were characterized using SANS [232], as were poly(ethylene glycol)-modified liposomes, which were studied for drug encapsulation and delivery [233] and for mRNA delivery [234]. The interaction of a poly(ethylene glycol)-polymer hybrid with lipid bilayer vesicles was investigated with the use of SANS [235]. By using contrast matching, domain formation in vesicles containing polymers was observed [236]. The accumulation of styrene oligomers in lipid bilayers was studied as a model for how plastic wastes interact with cell membranes [237]. The interaction of polythiophenes with model cell membranes was studied, which revealed that they associate with different regions of the bilayer depending on the specific chemistry of the terminal group of the polymer alkyl side chain [238]. The interaction of polymers that had been developed to have antibacterial action with lipid bilayer vesicles was investigated to learn how the polymer incorporates into the bilayer while still being able to function [239]. SANS revealed the oligomerization state of a polymer that selectively conducts protons, similar to some proteins, when incorporated into vesicles [240]. The location of monomers within the structure of the bilayer was also determined with SANS [241].

5.4. SANS for Studying the Kinetics of Lipid Bilayer Membranes

The non-destructive nature of neutrons and the high flux of modern sources provide opportunities for studying kinetic processes in membranes. Such studies require contrast-matching methods and deuterium labeling. The movement of water across a model membrane over time was resolved with SANS [242]. SANS was also used to evaluate the transition of lipid nanodisks to unilamellar vesicles [243]. Simple curvature effects can facilitate lipid exchange [244], for example, and so can methanol [245]. The transport of cholesterol between POPC vesicles and between leaflets of a single vesicle was found by SANS to be slower than previously reported [246]. Later work revealed that cholesterol exchange happens more rapidly when unsaturated PC lipids are included in the vesicles [247], but it happens more slowly in POPS vesicles [248]. The kinetics of lipid exchange enabled by the SARS-CoV-2 fusion peptide was also probed using SANS [195]. The peptide-driven lipid exchange caused by melittin and alamethicin was also investigated [249]. The peptide indolicidin speeds up lipid exchange between vesicles while also making them grow [250], as do the antimicrobial peptides Aurein 1.2, LL-37, and Lacticin Q [251] Other peptides have a similar effect [252,253,254]. A series of peptides designed to scramble lipids in a bilayer was studied by Nakao and coworkers using SANS and contrast-matching methods [255]. By tracking the integrated intensity in a specific q-range, they were able to measure the rate of change in the neutron scattering contrast and therefore the lipid exchange. Examples of intensity decay curves are shown in Figure 10.

5.5. SANS of Intact Cell Membranes

SANS has also been used to study membranes in intact cells and organisms. One of the main advantages of using SANS to study living cells is the lack of radiation damage by neutrons that possess much lower energy than the X-ray photons used for SAXS experiments. However, applying contrast variation methods is more difficult due to the toxicity of 2 H 2 O. Deuterium incorporation into plants also remains challenging [256]. Achieving the selective deuteration of components within cells in order to provide strong contrast against the remainder of the cell is not trivial. The structural information about the biomembrane contained in the data is also partially obscured by the other structures in the cell, which have common length scales. In spite of the challenges, SANS is a powerful tool for studying intact, living biological membranes, and examples of this follow.
One example of where SANS has been used to study intact cells is in investigations of the thylakoid membranes, which are part of the photosynthetic apparatus of plants [257,258,259,260,261,262,263,264,265,266,267]. Thylakoid membranes organize into stacks, and thus they give rise to diffraction features in SANS patterns that can be resolved from the remainder of the signal of the leaf of the plant or cell that is being measured. SANS studies have been performed on cyanobacteria [257,259,260,263,264], diatoms [258], algae [261,267], and higher plants [262,265,266]. The impact of environmental stressors, such as salt [267], selenium [265] and herbicide [264], on the structure of the membranes has also been studied. Cyanobacteria with mutations in the photosynthetic apparatus were studied using SANS to reveal the impact of the mutations on the structure of the membranes [259,260]. The response of the membranes to light has received considerable attention in various studies [257,258,259,262,264,266]. Ünnep and coworkers studied dark-adapted M. deliciosa (split-leaf philodendron) leaves and how they respond to exposure to light [266]. They observed a clear change in the thylakoid membrane signal after 8 min of light exposure, followed by 30 min in the dark, as can be seen in Figure 11.
SANS measurements have revealed information about the structure of other cellular membranes as well. SANS revealed the repeat spacing of the membranes within mitochondria [268]. SANS with contrast variation and SAXS were used to develop a model of living E. coli cells, which provided information about the inner and outer membranes of the bacteria [269,270]. Nickels and coworkers used contrast-matching methods to study B. subtilis cellular membranes in vivo [271]. By feeding the bacteria with deuterated fatty acids, it was possible to measure the samples in a mixture of H 2 O and 2 H 2 O, which minimized the scattering from everything in the cell, except for the acyl chain region of the cell membrane. Doing so allowed the team to measure the thickness of the cellular membrane, as can be seen in Figure 12, as well as to observe domains within its lateral structure [271].

6. Conclusions

SANS is a powerful tool for studying the structure of lipid bilayer membranes, and it will remain important in the future. When enhanced with contrast variation methods and selective deuterium labeling, it is possible to obtain information at the nanoscale about multi-component lipid mixtures, which is difficult, if not impossible, to to attain with the use of other experimental techniques. In addition, there are many opportunities to use these approaches in order to gain the insight needed to understand the complexity of cellular membranes. In particular, details about lipid distribution in samples made of more than one lipid are readily accessible, and the intrinsic contrast between proteins and lipids provides opportunities to study their interactions. Another area where future opportunities exist is in using SANS with contrast variation methods to investigate the structures and interactions of proteins that have been reconstituted into vesicles. In particular, using SANS and contrast variation methods would be a powerful tool for observing the oligomerization states of proteins or the complex formation, in the case of two or more different proteins, in lipid bilayer vesicles. The more native-like environment of the vesicle would allow for a more physically reasonable association to take place. In spite of the challenges intrinsic to selective deuterium labeling in living cells, the work of Nickels and coworkers demonstrates that such experiments can provide truly unique information about living cell membranes [271], and additional developments in this area have considerable potential for scientific impact.
A great deal can be learned using more traditional SANS data analysis approaches, which are readily accessible to both expert and novice users of the technique. However, not all of the more advanced approaches are necessarily accessible to the novice or casual user. Opportunities to improve the availability of such data analysis approaches exist. Furthermore, the continued development of more advanced modeling methods, such as those leveraging MD simulations, has excellent potential for revealing additional details when used in the analysis and interpretation of SANS data. The additional development of methods for modeling laterally inhomogeneous structures and making them widely accessible would also benefit the existing and potential user community. As is always the case with SAS data, care would need to be taken in order to avoid an over-interpretation of the data.

Funding

This research used resources at the Spallation Neutron Source, a DOE Office of Science User Facility operated by the Oak Ridge National Laboratory. This research used resources of the Advanced Photon Source, a U.S. Department of Energy (DOE) Office of Science user facility operated by the DOE Office of Science by Argonne National Laboratory under Contract No. DE-AC02-06CH11357.

Institutional Review Board Statement

Not applicable.

Informed Consent Statement

Not applicable.

Data Availability Statement

Data available from the author on request.

Acknowledgments

The author would like to thank X. Zuo of Argonne National Laboratory for assistance with the SAXS experiments. The author would like to thank G. Nagy of Oak Ridge National Laboratory for his thoughtful comments on the manuscript.

Conflicts of Interest

The author declares no conflict of interest. The funding agency had no role in the design of the study; in the collection, analyses, or interpretation of data; in the writing of the manuscript, or in the decision to publish the results.

Abbreviations

The following abbreviations were used in this manuscript:
SANSSmall-angle neutron scattering
SAXSSmall-angle X-ray scattering
MDMolecular dynamics
QENSQuasi-elastic neutron scattering
NSENeutron spin echo spectroscopy
SLDNeutron scattering length density
HGLipid head group
DHPC1,2-dihexanoyl-sn-glycero-3-phosphocholine
DLPC1,2-dilauroyl-sn-glycero-3-phosphocholine
DMPC1,2-dimyristoyl-sn-glycero-3-phosphocholine
DPPC1,2-dipalmitoyl-sn-glycero-3-phosphocholine
DSPC1,2-distearoyl-sn-glycero-3-phosphocholine
DPhPC1,2-diphytanyl-sn-glycero-3-phosphocholine
DMPG1,2-dimyristoyl-sn-glycero-3-phospho-(1’-rac-glycerol) sodium salt
POPC1-palmitoyl-2-oleoyl-sn-glycero-3-phosphocholine
POPG1-palmitoyl-2-oleoyl-sn-glycero-3-phospho-(1’-rac-glycerol) sodium salt
DOPC1,2-dioleoyl-sn-glycero-3-phosphocholine
POPS1-palmitoyl-2-oleoyl-sn-glycero-3-phospho-L-serine sodium salt
DSPE1,2-distearoyl-sn-glycero-3-phosphoethanolamine
POPS1-stearoyl-2-oleoyl-sn-glycero-3-phospho-L-serine sodium salt
DMPA1,2-dimyristoyl-sn-glycero-3-phosphate sodium salt
PCPhosphatidylcholine lipid head group
PGPhosphatidylglycerol lipid head group
PEPhosphatidylethanolamine lipid head group
PSPhosphatidylserine lipid head group
PAPhosphatic acid lipid head group

References

  1. Leventis, P.A.; Grinstein, S. The Distribution and Function of Phosphatidylserine in Cellular Membranes. Annu. Rev. Biophys. 2010, 39, 407–427. [Google Scholar] [CrossRef] [PubMed]
  2. Clejan, S.; Krulwich, T.A.; Mondrus, K.R.; Seto-Young, D. Membrane Lipid Composition of Obligately and Facultatively Alkalophilic Strains of Bacillus spp. J. Bacteriol. 1986, 168, 334–340. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  3. Morein, S.; Andersson, A.S.; Rilfors, L.; Lindblom, G. Wild-yype Escherichia coli Cells Regulate the Membrane Lipid Composition in a “Window” between Gel and Non-lamellar Structures. J. Biol. Chem. 1996, 271, 6801–6809. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  4. Devaux, P.F. Static and Dynamic Lipid Asymmetry in Cell Membranes. Biochemistry 1991, 30, 1163–1173. [Google Scholar] [CrossRef]
  5. Nagao, K.; Kimura, Y.; Mastuo, M.; Ueda, K. Lipid outward translocation by ABC proteins. FEBS Lett. 2010, 584, 2717–2723. [Google Scholar] [CrossRef] [Green Version]
  6. Jacobson, K.; Mouritsen, O.G.; Anderson, R.G.W. Lipid rafts: At a crossroad between cell biology and physics. Nat. Cell Biol. 2007, 9, 7–14. [Google Scholar] [CrossRef]
  7. Brahms, S.; Brahms, J. Determination of Protein Secondary Structrue in Solution by Vacuum Ultraviolet Circular Dichroism. J. Mol. Biol. 1980, 138, 149–178. [Google Scholar] [CrossRef]
  8. Wu, Y.; Huang, H.W.; Olah, G.A. Method of Oriented Circular Dichroism. Biophys. J. 1990, 57, 797–806. [Google Scholar] [CrossRef] [Green Version]
  9. Castellani, F.; van Rossum, B.; Diehl, A.; Schubert, M.; Rehbein, K.; Oschkinat, H. Structure of a Protein Determined by Solid-State Magic-Angle-Spinning NMR Spectroscopy. Nature 2002, 420, 98–102. [Google Scholar] [CrossRef]
  10. Haris, P.I.; Chapman, D. The Conformational Analysis of Peptides using Fourier Transform IR Spectroscopy. Biopolymers 1995, 37, 251–263. [Google Scholar] [CrossRef]
  11. Chung, L.A.; Lear, J.D.; DeGrado, W.F. Fluorescence Studies of the Secondary Structure and Orientation of a Model Ion Channel Peptide in Phospholipid Vesicles. Biochemistry 1992, 31, 6608–6616. [Google Scholar] [CrossRef] [PubMed]
  12. Veatch, S.L.; Keller, S.L. Organization in Lipid Membranes Containing Cholesterol. Phys. Rev. Lett. 2002, 89, 268101. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  13. Korlach, J.; Schwille, P.; Webb, W.W.; Feigenson, G.W. Characterization of Lipid Bilayer Phases by Confocal Microscopy and Fluorescence Correlation Spectroscopy. Proc. Natl. Acad. Sci. USA 1999, 96, 8461–8466. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  14. Pedersen, S.; Jorgensen, K.; Baekmark, T.R.; Mouritsen, O.G. Indirect Evidence for Lipid Domain Formation in the Transition Region of Phospholipid Bilayers by Two-Probe Fluorescence Energy Transfer. Biophys. J. 1996, 71, 554–560. [Google Scholar] [CrossRef] [Green Version]
  15. Douliez, J.P.; Leonard, A.; Dufourc, E.J. Restatement of Order Parameters in Biomembranes—Calculation of C-C Bond Order Parameters from C-D Quadrupolar Splittings. Biophys. J. 1995, 68, 1727–1739. [Google Scholar] [CrossRef] [Green Version]
  16. Petrache, H.I.; Dodd, S.W.; Brown, M.F. Area per Lipid and Acyl Length Distributions in Fluid Phosphatidylcholines Determined by 2H NMR Spectroscopy. Biophys. J. 2000, 79, 3172–3192. [Google Scholar] [CrossRef] [Green Version]
  17. Franks, N.P. Structural Analysis of Hydrated Egg Lecithin and Cholesterol Bilayers. 1. X-ray Diffraction. J. Mol. Biol. 1976, 100, 345–358. [Google Scholar] [CrossRef]
  18. Worcester, D.L.; Franks, N.P. Structural Analysis of Hydrated Egg Lecithin and Cholesterol Bilayers. 2. Neutron Diffraction. J. Mol. Biol. 1976, 100, 359–378. [Google Scholar] [CrossRef]
  19. Nagle, J.F.; Wilkinson, D.A. Lecithin Bilayers—Density Measurements and Molecular Interactions. Biophys. J. 1978, 23, 159–175. [Google Scholar] [CrossRef] [Green Version]
  20. Büldt, G.; Gally, H.U.; Seelig, J.; Zaccai, G. Neutron Diffraction Studies on Phosphatidylcholine Model Membranes. 1. Head Group Conformation. J. Mol. Biol. 1979, 134, 673–691. [Google Scholar] [CrossRef]
  21. Zaccai, G.; Buldt, G.; Seelig, A.; Seelig, J. Neutron Diffraction Studies on Phosphatidylcholine Model Membranes. 2. Chain Conformation and Segmental Disorder. J. Mol. Biol. 1979, 134, 693–706. [Google Scholar] [CrossRef]
  22. Wiener, M.C.; White, S.H. Fluid Bilayer Structure Determined by the Combined use of X-ray and Neutron Diffraction. 1. Fluid Bilayer Models and the Limits of Resolution. Biophys. J. 1991, 59, 162–173. [Google Scholar] [CrossRef] [Green Version]
  23. Wiener, M.C.; White, S.H. Fluid Bilayer Structure Determination by the Combined use of X-ray and Neutron Diffraction. 2. Composition-Space Refinement Method. Biophys. J. 1991, 59, 174–185. [Google Scholar] [CrossRef] [Green Version]
  24. Olah, G.A.; Huang, H.W.; Liu, W.H.; Wu, Y.L. Location of Ion-Binding Sites in the Gramicidin Channel by X-ray Diffraction. J. Mol. Biol. 1991, 218, 847–858. [Google Scholar] [CrossRef]
  25. Wiener, M.C.; White, S.H. Structure of a Fluid Dioleoylphosphatidylcholine Bilayer Determined by Joint Refinement of X-ray and Neutron Diffraction Data. 3. Complete Structure. Biophys. J. 1992, 61, 434–447. [Google Scholar] [CrossRef] [Green Version]
  26. Bunjes, N.; Schmidt, E.K.; Jonczyk, A.; Rippmann, F.; Beyer, D.; Ringsdorf, H.; Graber, P.; Knoll, W.; Naumann, R. Thiopeptide-Supported Lipid Layers on Solid Substrates. Langmuir 1997, 13, 6188–6194. [Google Scholar] [CrossRef]
  27. Majewski, J.; Wong, J.Y.; Park, C.K.; Seitz, M.; Israelachvili, J.N.; Smith, G.S. Structural Studies of Polymer-Cushioned Lipid Bilayers. Biophys. J. 1998, 75, 2363–2367. [Google Scholar] [CrossRef] [Green Version]
  28. Wong, J.Y.; Majewski, J.; Seitz, M.; Park, C.K.; Israelachvili, J.N.; Smith, G.S. Polymer-Cushioned Bilayers. I. A Structural Study of Various Preparation Methods using Neutron Reflectometry. Biophys. J. 1999, 77, 1445–1457. [Google Scholar] [CrossRef] [Green Version]
  29. Krueger, S.; Meuse, C.W.; Majkrzak, C.F.; Dura, J.A.; Berk, N.F.; Tarek, M.; Plant, A.L. Investigation of Hybrid Bilayer Membranes with Neutron Reflectometry: Probing the Interactions of Melittin. Langmuir 2001, 17, 511–521. [Google Scholar] [CrossRef]
  30. Salditt, T.; Munster, C.; Mennicke, U.; Ollinger, C.; Fragneto, G. Thermal Fluctuations of Oriented Lipid Membranes by Nonspecular Neutron Reflectometry. Langmuir 2003, 19, 7703–7711. [Google Scholar] [CrossRef]
  31. Perez-Salas, U.A.; Faucher, K.M.; Majkrzak, C.F.; Berk, N.F.; Krueger, S.; Chaikof, E.L. Characterization of a Biomimetic Polymeric Lipid Bilayer by Phase Sensitive Neutron Reflectometry. Langmuir 2003, 19, 7688–7694. [Google Scholar] [CrossRef]
  32. de Santis, A.; Scoppola, E.; Ottaviani, M.F.; Koutsioubas, A.; Barnsley, L.C.; Paduano, L.; D’Errico, G.; Russo Krauss, I. Order vs. Disorder: Cholesterol and Omega-3 Phospholipids Determine Biomembrane Organization. Int. J. Mol. Sci. 2022, 23, 5322. [Google Scholar] [CrossRef] [PubMed]
  33. Blaurock, A.E. Structure of Nerve Myelin Membrane—Proof of Low-Resolution Profile. J. Mol. Biol. 1971, 56, 35–52. [Google Scholar] [CrossRef] [Green Version]
  34. Hønger, T.; Mortensen, K.; Ipsen, J.H.; Lemmich, J.; Bauer, R.; Mouritsen, O.G. Anomalous Swelling of Multilamellar Lipid Bilayers in the Transition Region by Renormalization of Curvature Elasticity. Phys. Rev. Lett. 1994, 72, 3911–3914. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. Lemmich, J.; Mortensen, K.; Ipsen, J.H.; Honger, T.; Bauer, R.; Mouritsen, O.G. Small-Angle Neutron Scattering from Multilamellar Lipid Bilayers: Theory, Model, and Experiment. Phys. Rev. E 1996, 53, 5169–5180. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  36. Pabst, G.; Rappolt, M.; Amenitsch, H.; Laggner, P. Structural Information from Multilamellar Liposomes at Full Hydration: Full Q-Range Fitting with High Quality X-ray Data. Phys. Rev. E 2000, 62, 4000–4009. [Google Scholar] [CrossRef] [Green Version]
  37. Winter, I.; Pabst, G.; Rappolt, M.; Lohner, K. Refined Structure of 1,2-diacyl-P-O-ethylphosphatidylcholine Bilayer Membranes. Chem. Phys. Lipids 2001, 112, 137–150. [Google Scholar] [CrossRef]
  38. Hjelm, R.P. The Resolution of TOF Low-Q Diffractometers—Instrumental, Data Acquisition and Reduction Factors. J. Appl. Crystallogr. 1988, 21, 618–628. [Google Scholar] [CrossRef]
  39. Reeves, J.P.; Dowben, R.M. Formation and Properties of Thin-Walled Phospholipid Vesicles. J. Cell. Physiol. 1969, 73, 49–60. [Google Scholar] [CrossRef]
  40. Needham, D.; Evans, E. Structure and Mechanical Properties of Giant Lipid (DMPC) Vesicle Bilayers from 20.degree.C below to 10.degree.C above the Liquid Crystalline Phase Transition at 24.Degree.C. Biochemistry 1988, 27, 8261–8269. [Google Scholar] [CrossRef]
  41. Hope, M.J.; Bally, M.B.; Webb, G.; Cullis, P.R. Production of Large Unilamellar Vesicles by a Rapid Extrusion Procedure—Characterization of Size Distribution, Trapped Volume and Ability to Maintain a Membrane Potential. Biochim. Biophys. Acta 1985, 812, 55–65. [Google Scholar] [CrossRef]
  42. MacDonald, R.C.; MacDonald, R.I.; Menco, B.P.M.; Takeshita, K.; Subbarao, N.K.; Hu, L.R. Small-Volume Extrusion Apparatus for Preparation of Large, Unilamellar Vesicles. Biochim. Biophys. Acta 1991, 1061, 297–303. [Google Scholar] [CrossRef]
  43. Maulucci, G.; De Spirito, M.; Arcovito, G.; Boffi, F.; Castellano, A.C.; Briganti, G. Particle size distribution in DMPC vesicles solutions undergoing different sonication times. Biophys. J. 2005, 88, 3545–3550. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  44. Silva, R.; Ferreira, H.; Little, C.; Cavaco-Paulo, A. Effect of ultrasound parameters for unilamellar liposome preparation. Ultrason. Sonochem. 2010, 17, 628–632. [Google Scholar] [CrossRef] [Green Version]
  45. Angelova, M.I.; Dimitrov, D.S. Liposome Electroformation. Faraday Discuss. 1986, 81, 303–311. [Google Scholar] [CrossRef]
  46. Watwe, R.M.; Bellare, J.R. Manufacture of Liposomes—A Review. Curr. Sci. 1995, 68, 715–724. [Google Scholar]
  47. Nele, V.; Holme, M.N.; Kauscher, U.; Thomas, M.R.; Doutch, J.J.; Stevens, M.M. Effect of Formulation Method, Lipid Composition, and PEGylation on Vesicle Lamellarity: A Small-Angle Neutron Scattering Study. Langmuir 2019, 35, 6064–6074. [Google Scholar] [CrossRef] [Green Version]
  48. Scott, H.L.; Skinkle, A.; Kelley, E.G.; Waxham, M.N.; Levental, I.; Heberle, F.A. On the Mechanism of Bilayer Separation by Extrusion, or Why Your LUVs Are Not Really Unilamellar. Biophys. J. 2019, 117, 1381–1386. [Google Scholar] [CrossRef]
  49. Debye, P.; Bueche, A. Scattering by an Inhomogeneous Solid. J. Appl. Phys. 1949, 20, 518–525. [Google Scholar] [CrossRef]
  50. Guinier, A.; Fournet, G. Small-Angle Scattering of X-rays; Structure of Matter Series; John Wiley & Sons: New York, NY, USA, 1955. [Google Scholar]
  51. Pedersen, J.S. Analysis of small-angle scattering data from colloids and polymer solutions: Modeling and least-squares fitting. Adv. Colloid Interface Sci. 1997, 70, 171–210. [Google Scholar] [CrossRef]
  52. Sears, V.F. Neutron scattering lengths and cross sections. Neutron News 1992, 3, 26–37. [Google Scholar] [CrossRef]
  53. Heller, W.T. Small-angle neutron scattering and contrast variation: A powerful combination for studying biological structures. Acta Crystallogr. Sect. D 2010, 66, 1213–1217. [Google Scholar] [CrossRef] [PubMed]
  54. Elder, M.; Hitchcock, P.; Mason, R.; Shipley, G.G. A Refinement Analysis of the Crystallography of the Phospholipid, 1,2-Dilauroyl-DL-Phosphatidylethanolamine, and Some Remarks on Lipid-Lipid and Lipid-Protein Interactions. Proc. R. Soc. A Math. Phys. Eng. Sci. 1977, 354, 157–170. [Google Scholar] [CrossRef]
  55. Pearson, R.H.; Pascher, I. The molecular structure of lecithin dihydrate. Nature 1979, 281, 499–501. [Google Scholar] [CrossRef]
  56. Epand, R.M.; Epand, R.F. Studies of thermotropic phospholipid phase transitions using scanning densitometry. Chem. Phys. Lipids 1980, 27, 139–150. [Google Scholar] [CrossRef]
  57. Pascher, I.; Sundell, S.; Hauser, H. Polar group interaction and molecular packing of membrane lipids. The crystal structure of lysophosphatidylethanolamine. J. Mol. Biol. 1981, 153, 807–824. [Google Scholar] [CrossRef]
  58. Harlos, K.; Eibl, H.; Pascher, I.; Sundell, S. Conformation and packing properties of phosphatidic acid: The crystal structure of monosodium dimyristoylphosphatidate. Chem. Phys. Lipids 1984, 34, 115–126. [Google Scholar] [CrossRef]
  59. Seddon, J.M.; Cevc, G.; Kaye, R.D.; Marsh, D. X-ray Diffraction Study of the Polymorphism of Hydrated Diacyl- and Dialkylphosphatidylethanolamines. Biochemistry 1984, 23, 2634–2644. [Google Scholar] [CrossRef] [PubMed]
  60. Schmidt, G.; Knoll, W. Densitometric characterization of aqueous lipid dispersions. Berichte Der Bunsengesellschaft/Phys. Chem. Chem. Phys. 1985, 89, 36–43. [Google Scholar] [CrossRef]
  61. Marra, J. Direct measurement of the interaction between phosphatidylglycerol bilayers in aqueous electrolyte solutions. Biophys. J. 1986, 50, 815–825. [Google Scholar] [CrossRef] [Green Version]
  62. McIntosh, T.J.; Simon, S.A. Area per Molecule and Distribution of Water in Fully Hydrated Dilauroylphosphatidylethanolamine Bilayers. Biochemistry 1986, 25, 4948–4952. [Google Scholar] [CrossRef] [PubMed]
  63. Laggner, P.; Lohner, K.; Degovics, G.; Müller, K.; Schuster, A. Structure and thermodynamics of the dihexadecylphosphatidylcholine-Water system. Chem. Phys. Lipids 1987, 44, 31–60. [Google Scholar] [CrossRef]
  64. Pascher, I.; Sundell, S.; Harlos, K.; Eibl, H. Conformation and packing properties of membrane lipids: The crystal structure of sodium dimyristoylphosphatidylglycerol. Biochim. Biophys. Acta (BBA)-Biomembr. 1987, 896, 77–88. [Google Scholar] [CrossRef]
  65. Wilkinson, D.A.; Tirrell, D.A.; Turek, A.B.; McIntosh, T.J. Tris buffer causes acyl chain interdigitation in phosphatidylglycerol. Biochim. Biophys. Acta (BBA)-Biomembr. 1987, 905, 447–453. [Google Scholar] [CrossRef]
  66. Nagle, J.F.; Wiener, M.C. Structure of fully hydrated bilayer dispersions. Biochim. Biophys. Acta (BBA)-Biomembr. 1988, 942, 1–10. [Google Scholar] [CrossRef]
  67. Koynova, R.; Hinz, H.J. Metastable behaviour of saturated phosphatidylethanolamines: A densitometric study. Chem. Phys. Lipids 1990, 54, 67–72. [Google Scholar] [CrossRef]
  68. Pascher, I.; Lundmark, M.; Nyholm, P.G.; Sundell, S. Crystal structures of membrane lipids. Biochim. Biophys. Acta (BBA)-Rev. Biomembr. 1992, 1113, 339–373. [Google Scholar] [CrossRef]
  69. Wiener, M.C.; White, S.H. Structure of a fluid dioleoylphosphatidylcholine bilayer determined by joint refinement of x-ray and neutron diffraction data. II. Distribution and packing of terminal methyl groups. Biophys. J. 1992, 61, 428–433. [Google Scholar] [CrossRef] [Green Version]
  70. Sun, W.J.; Suter, R.M.; Knewtson, M.A.; Worthington, C.R.; Tristram-Nagle, S.; Zhang, R.; Nagle, J.F. Order and disorder in fully hydrated unoriented bilayers of gel-phase dipalmitoylphosphatidylcholine. Phys. Rev. E 1994, 49, 4665–4676. [Google Scholar] [CrossRef]
  71. Armen, R.S.; Uitto, O.D.; Feller, S.E. Phospholipid Component Volumes: Determination and Application to Bilayer Structure Calculations. Biophys. J. 1998, 75, 734–744. [Google Scholar] [CrossRef] [Green Version]
  72. Petrache, H.I.; Tristram-Nagle, S.; Gawrisch, K.; Harries, D.; Parsegian, V.A.; Nagle, J.F. Structure and Fluctuations of Charged Phosphatidylserine Bilayers in the Absence of Salt. Biophys. J. 2004, 86, 1574–1586. [Google Scholar] [CrossRef]
  73. Kučerka, N.; Liu, Y.; Chu, N.; Petrache, H.I.; Tristram-Nagle, S.; Nagle, J.F. Structure of fully hydrated fluid phase DMPC and DLPC lipid bilayers using X-ray scattering from oriented multilamellar arrays and from unilamellar vesicles. Biophys. J. 2005, 88, 2626–2637. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  74. Kučerka, N.; Tristram-Nagle, S.; Nagle, J.F. Structure of Fully Hydrated Fluid Phase Lipid Bilayers with Monounsaturated Chains. J. Membr. Biol. 2006, 208, 193–202. [Google Scholar] [CrossRef] [PubMed]
  75. Kučerka, N.; Tristram-Nagle, S.; Nagle, J.F. Closer look at structure of fully hydrated fluid phase DPPC bilayers. Biophys. J. 2006, 90, L83–L85. [Google Scholar] [CrossRef] [Green Version]
  76. Uhríková, D.; Rybár, P.; Hianik, T.; Balgavý, P. Component volumes of unsaturated phosphatidylcholines in fluid bilayers: A densitometric study. Chem. Phys. Lipids 2007, 145, 97–105. [Google Scholar] [CrossRef] [PubMed]
  77. Greenwood, A.I.; Pan, J.; Mills, T.T.; Nagle, J.F.; Epand, R.M.; Tristram-Nagle, S. CRAC motif peptide of the HIV-1 gp41 protein thins SOPC membranes and interacts with cholesterol. Biochim. Biophys. Acta-Biomembr. 2008, 1778, 1120–1130. [Google Scholar] [CrossRef] [Green Version]
  78. Pabst, G.; Grage, S.L.; Danner-Pongratz, S.; Jing, W.; Ulrich, A.S.; Watts, A.; Lohner, K.; Hickel, A. Membrane thickening by the antimicrobial peptide PGLa. Biophys. J. 2008, 95, 5779–5788. [Google Scholar] [CrossRef] [Green Version]
  79. Guler, S.D.; Ghosh, D.D.; Pan, J.; Mathai, J.C.; Zeidel, M.L.; Nagle, J.F.; Tristram-Nagle, S. Effects of ether vs. ester linkage on lipid bilayer structure and water permeability. Chem. Phys. Lipids 2009, 160, 33–44. [Google Scholar] [CrossRef] [Green Version]
  80. Marsh, D. Molecular volumes of phospholipids and glycolipids in membranes. Chem. Phys. Lipids 2010, 163, 667–677. [Google Scholar] [CrossRef] [Green Version]
  81. Tristram-Nagle, S.; Kim, D.J.; Akhunzada, N.; Kučerka, N.; Mathai, J.C.; Katsaras, J.; Zeidel, M.; Nagle, J.F. Structure and water permeability of fully hydrated diphytanoylPC. Chem. Phys. Lipids 2010, 163, 630–637. [Google Scholar] [CrossRef] [Green Version]
  82. Pan, J.; Heberle, F.A.; Tristram-Nagle, S.; Szymanski, M.; Koepfinger, M.; Katsaras, J.; Kučerka, N. Molecular structures of fluid phase phosphatidylglycerol bilayers as determined by small angle neutron and X-ray scattering. Biochim. Biophys. Acta-Biomembr. 2012, 1818, 2135–2148. [Google Scholar] [CrossRef]
  83. Kučerka, N.; van Oosten, B.; Pan, J.; Heberle, F.A.; Harroun, T.A.; Katsaras, J. Molecular Structures of Fluid Phosphatidylethanolamine Bilayers Obtained from Simulation-to-Experiment Comparisons and Experimental Scattering Density Profiles. J. Phys. Chem. B 2015, 119, 1947–1956. [Google Scholar] [CrossRef] [Green Version]
  84. Murugova, T.; Ivankov, O.; Ermakova, E.; Kondela, T.; Hrubovčák, P.; Skoi, V.; Kuklin, A.; Kučerka, N. Structural changes introduced by cholesterol and melatonin to the model membranes mimicking preclinical conformational diseases. Gen. Physiol. Biophys. 2020, 39, 135–144. [Google Scholar] [CrossRef] [PubMed]
  85. Jacrot, B. The study of biological structures by neutron scattering from solution. Rep. Prog. Phys. 1976, 39, 911–953. [Google Scholar] [CrossRef]
  86. Breyton, C.; Gabel, F.; Lethier, M.; Flayhan, A.; Durand, G.; Jault, J.M.; Juillan-Binard, C.; Imbert, L.; Moulin, M.; Ravaud, S.; et al. Small Angle Neutron Scattering for the Study of Solubilised Membrane Proteins. Eur. Phys. J. E 2013, 36, 71. [Google Scholar] [CrossRef]
  87. Kienzle, P. Neutron Activation and Scattering Calculator. Available online: https://www.ncnr.nist.gov/resources/activation/ (accessed on 18 May 2022).
  88. Heller, W.T.; Cuneo, M.; Debeer-Schmitt, L.; Do, C.; He, L.; Heroux, L.; Littrell, K.; Pingali, S.V.; Qian, S.; Stanley, C.; et al. The Suite of Small-Angle Neutron Scattering Instruments at Oak Ridge National Laboratory. J. Appl. Crystallogr. 2018, 51, 242–248. [Google Scholar] [CrossRef]
  89. Heller, W.T.; Zolnierczuk, P.A. The Helix-to-Sheet Transition of an HIV-1 Fusion Peptide Derivative Changes the Mechanical Properties of Lipid Bilayer Membranes. Biochim. Biophys. Acta-Biomembr. 2019, 1861, 565–572. [Google Scholar] [CrossRef] [PubMed]
  90. Sadler, D.M.; Reiss-Husson, F.; Rivas, E. Thickness measurements of single walled dimyristoyl phosphatidylcholine vesicles by neutron scattering. Chem. Phys. Lipids 1990, 52, 41–48. [Google Scholar] [CrossRef]
  91. Dubničkova, M.; Kiselev, M.; Kutuzov, S.; Devínsky, F.; Gordeliy, V.; Balgavý, P. Effect of N-lauryl-N,N-dimethylamine N-oxide on dimyristoyl phosphatidylcholine bilayer thickness: A small-angle neutron scattering study. Gen. Physiol. Biophys. 1997, 16, 175–188. [Google Scholar]
  92. Shashkov, S.N.; Kiselev, M.A.; Tioutiounnikov, S.N.; Kiselev, A.M.; Lesieur, P. The study of DMSO/water and DPPC/DMSO/water system by means of the X-ray, neutron small-angle scattering, calorimetry and IR spectroscopy. Phys. B Condens. Matter 1999, 271, 184–191. [Google Scholar] [CrossRef]
  93. Gutberlet, T.; Kiselev, M.; Heerklotz, H.; Klose, G. SANS study of mixed POPC/C12En aggregates. Phys. B Condens. Matter 2000, 276–278, 381–383. [Google Scholar] [CrossRef]
  94. Kučerka, N.; Uhríková, D.; Teixeira, J.; Balgavý, P. Bilayer thickness in unilamellar phosphatidylcholine vesicles: Small-angle neutron scattering using contrast variation. Phys. B Condens. Matter 2004, 350, E639–E642. [Google Scholar] [CrossRef]
  95. Gordeliy, V.I.; Cherezov, V.; Teixeira, J. Strength of thermal undulations of phospholipid membranes. Phys. Rev. E 2005, 72, 061913. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  96. Hirai, M.; Hirai, H.; Koizumi, M.; Kasahara, K.; Yuyama, K.; Suzuki, N. Structure of raft-model membrane by using the inverse contrast variation neutron scattering method. Phys. B Condens. Matter 2006, 385-386, 868–870. [Google Scholar] [CrossRef]
  97. Schrödinger, LLC. The PyMOL Molecular Graphics System, Version 2.5.2; Schrödinger, LLC: New York, NY, USA, 2021. [Google Scholar]
  98. Doucet, M.; Cho, J.H.; Alina, G.; Bakker, J.; Bouwman, W.; Butler, P.; Campbell, K.; Gonzales, M.; Heenan, R.; Jackson, A.; et al. SasView Version 5.0. Available online: https://www.sasview.org/ (accessed on 22 September 2022).
  99. Qian, S.; Heller, W.T. Peptide-Induced Asymmetric Distribution of Charged Lipids in a Vesicle Bilayer Revealed by Small-Angle Neutron Scattering. J. Phys. Chem. B 2011, 115, 9831–9837. [Google Scholar] [CrossRef]
  100. Qian, S.; Rai, D.; Heller, W.T. Alamethicin Disrupts the Cholesterol Distribution in Dimyristoyl Phosphatidylcholine-Cholesterol Lipid Bilayers. J. Phys. Chem. B 2014, 118, 11200–11208. [Google Scholar] [CrossRef]
  101. Rai, D.K.; Qian, S.; Heller, W.T. The Interaction of Melittin with Dimyristoyl Phosphatidylcholine-Dimyristoyl Phosphatidylserine Lipid Bilayer Membranes. Biochim. Biophys. Acta-Biomembr. 2016, 1858, 2788–2794. [Google Scholar] [CrossRef]
  102. Heller, W.T. A Small-Angle Neutron Scattering Study of the Physical Mechanism that Drives the Action of a Viral Fusion Peptide. Chem. Phys. Lipids 2021, 234, 105022. [Google Scholar] [CrossRef]
  103. Kučerka, N.; Nagle, J.F.; Feller, S.E.; Balgavý, P. Models to analyze small-angle neutron scattering from unilamellar lipid vesicles. Phys. Rev. E 2004, 69, 051903. [Google Scholar] [CrossRef] [Green Version]
  104. Kučerka, N.; Kiselev, M.A.; Balgavý, P. Determination of bilayer thickness and lipid surface area in unilamellar dimyristoylphosphatidylcholine vesicles from small-angle neutron scattering curves: A comparison of evaluation methods. Eur. Biophys. J. Biophys. Lett. 2004, 33, 328–334. [Google Scholar] [CrossRef]
  105. Belička, M.; Kučerka, N.; Uhríková, D.; Islamov, A.K.; Kuklin, A.I.; Devínsky, F.; Balgavý, P. Effects of N,N-dimethyl-N-alkylamine-N-oxides on DOPC bilayers in unilamellar vesicles: Small-angle neutron scattering study. Eur. Biophys. J. Biophys. Lett. 2014, 43, 179–189. [Google Scholar] [CrossRef] [PubMed]
  106. Tan, L.; Elkins, J.G.; Davison, B.H.; Kelley, E.G.; Nickels, J. Implementation of a self-consistent slab model of bilayer structure in the SasView Suite. J. Appl. Crystallogr. 2021, 54, 363–370. [Google Scholar] [CrossRef]
  107. Nagle, J.F.; Tristram-Nagle, S. Structure of Lipid Bilayers. Biochim. Biophys. Acta (BBA)-Rev. Biomembr. 2000, 1469, 159–195. [Google Scholar] [CrossRef] [Green Version]
  108. Kučerka, N.; Holland, B.W.; Gray, C.G.; Tomberli, B.; Katsaras, J. Scattering Density Profile Model of POPG Bilayers As Determined by Molecular Dynamics Simulations and Small-Angle Neutron and X-ray Scattering Experiments. J. Phys. Chem. B 2012, 116, 232–239. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  109. Kučerka, N.; Nagle, J.F.; Sachs, J.N.; Feller, S.E.; Pencer, J.; Jackson, A.; Katsaras, J. Lipid Bilayer Structure Determined by the Simultaneous Analysis of Neutron and X-Ray Scattering Data. Biophys. J. 2008, 95, 2356–2367. [Google Scholar] [CrossRef] [Green Version]
  110. Klauda, J.B.; Kučerka, N.; Brooks, B.R.; Pastor, R.W.; Nagle, J.F. Simulation-Based Methods for Interpreting X-ray Data from Lipid Bilayers. Biophys. J. 2006, 90, 2796–2807. [Google Scholar] [CrossRef] [Green Version]
  111. Fogarty, J.C.; Arjunwadkar, M.; Pandit, S.A.; Pan, J. Atomically detailed lipid bilayer models for the interpretation of small angle neutron and X-ray scattering data. Biochim. Biophys. Acta-Biomembr. 2015, 1848, 662–672. [Google Scholar] [CrossRef] [Green Version]
  112. Lewis-Laurent, A.; Doktorova, M.; Heberle, F.A.; Marquardt, D. Vesicle Viewer: Online visualization and analysis of small-angle scattering from lipid vesicles. Biophys. J. 2021, 120, 4639–4648. [Google Scholar] [CrossRef]
  113. Róg, T.; Murzyn, K.; Hinsen, K.; Kneller, G.R. nMoldyn: A program package for a neutron scattering oriented analysis of molecular dynamics simulations. J. Comput. Chem. 2003, 24, 657–667. [Google Scholar] [CrossRef]
  114. Lindner, B.; Smith, J.C. Sassena—X-ray and neutron scattering calculated from molecular dynamics trajectories using massively parallel computers. Comput. Phys. Commun. 2012, 183, 1491–1501. [Google Scholar] [CrossRef]
  115. Feller, S.E.; Yin, D.; Pastor, R.W.; MacKerell, A.D. Molecular dynamics simulation of unsaturated lipid bilayers at low hydration: Parameterization and comparison with diffraction studies. Biophys. J. 1997, 73, 2269–2279. [Google Scholar] [CrossRef] [Green Version]
  116. Sachs, J.N.; Petrache, H.I.; Woolf, T.B. Interpretation of small angle X-ray measurements guided by molecular dynamics simulations of lipid bilayers. Chem. Phys. Lipids 2003, 126, 211–223. [Google Scholar] [CrossRef] [PubMed]
  117. Benz, R.W.; Castro-Román, F.; Tobias, D.J.; White, S.H. Experimental Validation of Molecular Dynamics Simulations of Lipid Bilayers: A New Approach. Biophys. J. 2005, 88, 805–817. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  118. Kučerka, N.; Katsaras, J.; Nagle, J.F. Comparing membrane simulations to scattering experiments: Introducing the SIMtoEXP software. J. Membr. Biol. 2010, 235, 43–50. [Google Scholar] [CrossRef] [Green Version]
  119. Kučerka, N.; Nieh, M.P.; Katsaras, J. Fluid phase lipid areas and bilayer thicknesses of commonly used phosphatidylcholines as a function of temperature. Biochim. Biophys. Acta-Biomembr. 2011, 1808, 2761–2771. [Google Scholar] [CrossRef]
  120. Pan, J.; Cheng, X.; Monticelli, L.; Heberle, F.A.; Kučerka, N.; Tieleman, D.P.; Katsaras, J. The molecular structure of a phosphatidylserine bilayer determined by scattering and molecular dynamics simulations. Soft Matter 2014, 10, 3716–3725. [Google Scholar] [CrossRef]
  121. Pan, J.; Marquardt, D.; Heberle, F.A.; Kučerka, N.; Katsaras, J. Revisiting the bilayer structures of fluid phase phosphatidylglycerol lipids: Accounting for exchangeable hydrogens. Biochim. Biophys. Acta-Biomembr. 2014, 1838, 2966–2969. [Google Scholar] [CrossRef] [Green Version]
  122. Marrink, S.J.; Risselada, H.J.; Yefimov, S.; Tieleman, D.P.; de Vries, A.H. The MARTINI Force Field: Coarse Grained Model for Biomolecular Simulations. J. Phys. Chem. B 2007, 111, 7812–7824. [Google Scholar] [CrossRef] [Green Version]
  123. Carrillo, J.M.Y.; Katsaras, J.; Sumpter, B.G.; Ashkar, R. A Computational Approach for Modeling Neutron Scattering Data from Lipid Bilayers. J. Chem. Theory Comput. 2017, 13, 916–925. [Google Scholar] [CrossRef]
  124. Qian, S.; Heller, W.T. Melittin-Induced Cholesterol Reorganization in Lipid Bilayer Membranes. Biochim. Biophys. Acta-Biomembr. 2015, 1848, 2253–2260. [Google Scholar] [CrossRef] [Green Version]
  125. Heller, W.T.; Rai, D.K. Changes in Lipid Bilayer Structure Caused by the Helix-to-Sheet Transition of an HIV-1 GP41 Fusion Peptide Derivative. Chem. Phys. Lipids 2017, 203, 46–53. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  126. Heberle, F.A.; Petruzielo, R.S.; Pan, J.; Drazba, P.; Kučerka, N.; Standaert, R.F.; Feigenson, G.W.; Katsaras, J. Bilayer Thickness Mismatch Controls Domain Size in Model Membranes. J. Am. Chem. Soc. 2013, 135, 6853–6859. [Google Scholar] [CrossRef] [PubMed]
  127. Moore, P.B. Small-Angle Scattering—Information Content and Error Analysis. J. Appl. Crystallogr. 1980, 13, 168–175. [Google Scholar] [CrossRef]
  128. Svergun, D.I.; Volkov, V.V.; Kozin, M.B.; Stuhrmann, H.B. New developments in direct shape determination from small-angle scattering. 2. Uniqueness. Acta Crystallogr. Sect. A 1996, 52, 419–426. [Google Scholar] [CrossRef] [Green Version]
  129. Balgavý, P.; Dubnickova, M.; Kučerka, N.; Kiselev, M.A.; Yaradaikin, S.P.; Uhríková, D. Bilayer thickness and lipid interface area in unilamellar extruded 1,2-diacylphosphatidylcholine liposomes: A small-angle neutron scattering study. Biochim. Biophys. Acta-Biomembr. 2001, 1512, 40–52. [Google Scholar] [CrossRef] [Green Version]
  130. Mason, P.C.; Gaulin, B.D.; Epand, R.M.; Katsaras, J. Critical swelling in single phospholipid bilayers. Phys. Rev. E 2000, 61, 5634–5639. [Google Scholar] [CrossRef]
  131. Kuklin, A.; Zabelskii, D.; Gordeliy, I.; Teixeira, J.; Brûlet, A.; Chupin, V.; Cherezov, V.; Gordeliy, V. On the Origin of the Anomalous Behavior of Lipid Membrane Properties in the Vicinity of the Chain-Melting Phase Transition. Sci. Rep. 2020, 10, 5749. [Google Scholar] [CrossRef] [Green Version]
  132. Mason, P.C.; Gaulin, B.D.; Epand, R.M.; Wignall, G.D.; Lin, J.S. Small angle neutron scattering and calorimetric studies of large unilamellar vesicles of the phospholipid dipalmitoylphosphatidylcholine. Phys. Rev. E 1999, 59, 3361–3367. [Google Scholar] [CrossRef]
  133. Kiselev, M.A.; Zemlyanaya, E.V.; Aswal, V.K.; Neubert, R.H.H. What can we learn about the lipid vesicle structure from the small-angle neutron scattering experiment? Eur. Biophys. J. Biophys. Lett. 2006, 35, 477–493. [Google Scholar] [CrossRef]
  134. Pencer, J.; Krueger, S.; Adams, C.P.; Katsaras, J. Method of separated form factors for polydisperse vesicles. J. Appl. Crystallogr. 2006, 39, 293–303. [Google Scholar] [CrossRef]
  135. Uhríková, D.; Kučerka, N.; Teixeira, J.; Gordeliy, V.; Balgavý, P. Structural changes in dipalmitoylphosphatidylcholine bilayer promoted by Ca2+ ions: A small-angle neutron scattering study. Chem. Phys. Lipids 2008, 155, 80–89. [Google Scholar] [CrossRef] [PubMed]
  136. Meure, L.A.; Knott, R.; Foster, N.R.; Dehghani, F. The Depressurization of an Expanded Solution into Aqueous Media for the Bulk Production of Liposomes. Langmuir 2009, 25, 326–337. [Google Scholar] [CrossRef] [PubMed]
  137. Kučerka, N.; Gallová, J.; Uhríková, D.; Balgavý, P.; Bulacu, M.; Marrink, S.J.; Katsaras, J. Areas of Monounsaturated Diacylphosphatidylcholines. Biophys. J. 2009, 97, 1926–1932. [Google Scholar] [CrossRef] [Green Version]
  138. Marquardt, D.; Heberle, F.A.; Pan, J.; Cheng, X.; Pabst, G.; Harroun, T.A.; Kučerka, N.; Katsaras, J. The structures of polyunsaturated lipid bilayers by joint refinement of neutron and X-ray scattering data. Chem. Phys. Lipids 2020, 229, 104892. [Google Scholar] [CrossRef]
  139. Pan, J.; Cheng, X.; Sharp, M.; Ho, C.S.; Khadka, N.; Katsaras, J. Structural and mechanical properties of cardiolipin lipid bilayers determined using neutron spin echo, small angle neutron and X-ray scattering, and molecular dynamics simulations. Soft Matter 2015, 11, 130–138. [Google Scholar] [CrossRef]
  140. Doktorova, M.; Kučerka, N.; Kinnun, J.J.; Pan, J.; Marquardt, D.; Scott, H.L.; Venable, R.M.; Pastor, R.W.; Wassall, S.R.; Katsaras, J.; et al. Molecular Structure of Sphingomyelin in Fluid Phase Bilayers Determined by the Joint Analysis of Small-Angle Neutron and X-ray Scattering Data Published as part of The Journal of Physical Chemistry virtual special issue “Computational and Experimental Advances in Biomembranes”. J. Phys. Chem. B 2020, 124, 5186–5200. [Google Scholar] [CrossRef] [PubMed]
  141. Cantu, L.; Corti, M.; Del Favero, E.; Dubois, M.; Zemb, T.N. Combined small-angle X-ray and neutron scattering experiments for thickness characterization of ganglioside bilayers. J. Phys. Chem. B 1998, 102, 5737–5743. [Google Scholar] [CrossRef]
  142. D’Errico, G.; Silipo, A.; Mangiapia, G.; Molinaro, A.; Paduano, L.; Lanzetta, R. Mesoscopic and microstructural characterization of liposomes formed by the lipooligosaccharide from Salmonella minnesota strain 595 (Re mutant). Phys. Chem. Chem. Phys. 2009, 11, 2314–2322. [Google Scholar] [CrossRef]
  143. D’Errico, G.; Silipo, A.; Mangiapia, G.; Vitiello, G.; Radulescu, A.; Molinaro, A.; Lanzetta, R.; Paduano, L. Characterization of liposomes formed by lipopolysaccharides from Burkholderia cenocepacia, Burkholderia multivorans and Agrobacterium tumefaciens: From the molecular structure to the aggregate architecture. Phys. Chem. Chem. Phys. 2010, 12, 13574–13585. [Google Scholar] [CrossRef]
  144. Matviykiv, S.; Deyhle, H.; Kohlbrecher, J.; Neuhaus, F.; Zumbuehl, A.; Muller, B. Small-Angle Neutron Scattering Study of Temperature-Induced Structural Changes in Liposomes. Langmuir 2019, 35, 11210–11216. [Google Scholar] [CrossRef]
  145. Frampton, M.B.; Yakoub, D.; Katsaras, J.; Zelisko, P.M.; Marquardt, D. A calorimetric, volumetric and combined SANS and SAXS study of hybrid siloxane phosphocholine bilayers. Chem. Phys. Lipids 2021, 241, 105149. [Google Scholar] [CrossRef] [PubMed]
  146. Woodka, A.C.; Butler, P.D.; Porcar, L.; Farago, B.; Nagao, M. Lipid Bilayers and Membrane Dynamics: Insight into Thickness Fluctuations. Phys. Rev. Lett. 2012, 109, 058102. [Google Scholar] [CrossRef] [PubMed]
  147. Ashkar, R.; Nagao, M.; Butler, P.D.; Woodka, A.C.; Sen, M.K.; Koga, T. Tuning Membrane Thickness Fluctuations in Model Lipid Bilayers. Biophys. J. 2015, 109, 106–112. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  148. Nagao, M.; Kelley, E.G.; Ashkar, R.; Bradbury, R.; Butler, P.D. Probing Elastic and Viscous Properties of Phospholipid Bilayers Using Neutron Spin Echo Spectroscopy. J. Phys. Chem. Lett. 2017, 8, 4679–4684. [Google Scholar] [CrossRef] [PubMed]
  149. Gupta, S.; De Mel, J.U.; Perera, R.M.; Zolnierczuk, P.; Bleuel, M.; Faraone, A.; Schneider, G.J. Dynamics of Phospholipid Membranes beyond Thermal Undulations. J. Phys. Chem. Lett. 2018, 9, 2956–2960. [Google Scholar] [CrossRef]
  150. Kelley, E.G.; Butler, P.D.; Nagao, M. Scaling of lipid membrane rigidity with domain area fraction. Soft Matter 2019, 15, 2762–2767. [Google Scholar] [CrossRef]
  151. Chakraborty, S.; Doktorova, M.; Molugu, T.R.; Heberle, F.A.; Scott, H.L.; Dzikovski, B.; Nagao, M.; Stingaciu, L.R.; Standaert, R.F.; Barrera, F.N.; et al. How cholesterol stiffens unsaturated lipid membranes. Proc. Natl. Acad. Sci. USA 2020, 117, 21896–21905. [Google Scholar] [CrossRef]
  152. Kelley, E.G.; Butler, P.D.; Ashkar, R.; Bradbury, R.; Nagao, M. Scaling relationships for the elastic moduli and viscosity of mixed lipid membranes. Proc. Natl. Acad. Sci. USA 2020, 117, 23365–23373. [Google Scholar] [CrossRef]
  153. Kelley, E.G.; Nagao, M.; Butler, P.D.; Porcar, L.; Farago, B. Enhanced dynamics in the anomalous melting regime of DMPG lipid membranes. Struct. Dyn. 2020, 7, 054704. [Google Scholar] [CrossRef]
  154. Rickeard, B.W.; Nguyen, M.H.L.; DiPasquale, M.; Yip, C.G.; Baker, H.; Heberle, F.A.; Zuo, X.; Kelley, E.G.; Nagao, M.; Marquardt, D. Transverse lipid organization dictates bending fluctuations in model plasma membranes. Nanoscale 2020, 12, 1438–1447. [Google Scholar] [CrossRef]
  155. Gallova, J.; Uhríková, D.; Hanulova, M.; Teixeira, J.; Balgavý, P. Bilayer thickness in unilamellar extruded 1,2-dimyristoleoyl and 1,2-dierucoyl phosphatidylcholine vesicles: SANS contrast variation study of cholesterol effect. Colloids Surf. B-Biointerfaces 2004, 38, 11–14. [Google Scholar] [CrossRef] [PubMed]
  156. Gallova, J.; Uhríková, D.; Islamov, A.; Kuklin, A.; Balgavý, P. Effect of cholesterol on the bilayer thickness in unilamellar extruded DLPC and DOPC liposomes: SANS contrast variation study. Gen. Physiol. Biophys. 2004, 23, 113–128. [Google Scholar]
  157. Kučerka, N.; Pencer, J.; Nieh, M.P.; Katsaras, J. Influence of cholesterol on the bilayer properties of monounsaturated phosphatidylcholine unilamellar vesicles. Eur. Phys. J. E 2007, 23, 247–254. [Google Scholar] [CrossRef] [PubMed]
  158. Gallova, J.; Uhríková, D.; Kucerka, N.; Teixeira, J.; Balgavý, P. Hydrophobic thickness, lipid surface area and polar region hydration in monounsaturated diacylphosphatidylcholine bilayers: SANS study of effects of cholesterol and beta-sitosterol in unilamellar vesicles. Biochim. Biophys. Acta-Biomembr. 2008, 1778, 2627–2632. [Google Scholar] [CrossRef] [Green Version]
  159. Kucerka, N.; Nieh, M.P.; Katsaras, J. Asymmetric Distribution of Cholesterol in Unilamellar Vesicles of Monounsaturated Phospholipids. Langmuir 2009, 25, 13522–13527. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  160. Gallova, J.; Uhríková, D.; Kučerka, N.; Teixeira, J.; Balgavý, P. Partial area of cholesterol in monounsaturated diacylphosphatidylcholine bilayers. Chem. Phys. Lipids 2010, 163, 765–770. [Google Scholar] [CrossRef] [Green Version]
  161. Gallová, J.; Uhríková, D.; Kučerka, N.; Doktorovová, S.; Funari, S.S.; Teixeira, J.; Balgavý, P. The effects of cholesterol and beta-sitosterol on the structure of saturated diacylphosphatidylcholine bilayers. Eur. Biophys. J. Biophys. Lett. 2011, 40, 153–163. [Google Scholar] [CrossRef] [Green Version]
  162. Garg, S.; Castro-Roman, F.; Porcar, L.; Butler, P.; Bautista, P.J.; Krzyzanowski, N.; Perez-Salas, U. Cholesterol solubility limit in lipid membranes probed by small angle neutron scattering and MD simulations. Soft Matter 2014, 10, 9313–9317. [Google Scholar] [CrossRef] [Green Version]
  163. Pan, J.; Cheng, X.; Heberle, F.A.; Mostofian, B.; Kučerka, N.; Drazba, P.; Katsaras, J. Interactions between Ether Phospholipids and Cholesterol As Determined by Scattering and Molecular Dynamics Simulations. J. Phys. Chem. B 2012, 116, 14829–14838. [Google Scholar] [CrossRef] [Green Version]
  164. Zbytovska, J.; Kiselev, M.A.; Funari, S.S.; Garamus, V.M.; Wartewrig, S.; Palat, K.; Neubert, R. Influence of cholesterol on the structure of stratum corneum lipid model membrane. Colloids Surf. A-Physicochem. Eng. Asp. 2008, 328, 90–99. [Google Scholar] [CrossRef]
  165. Krivanek, R.; Jeworrek, C.; Czeslik, C.; Winter, R. Composition Fluctuations in Phospholipid-Sterol Vesicles—A Small-angle Neutron Scattering Study. Z. Fur Phys. Chem.-Int. J. Res. Phys. Chem. Chem. Phys. 2008, 222, 1679–1692. [Google Scholar] [CrossRef]
  166. Nieh, M.P.; Raghunathan, V.A.; Kline, S.R.; Harroun, T.A.; Huang, C.Y.; Pencer, J.; Katsaras, J. Spontaneously formed unilamellar vesicles with path-dependent size distribution. Langmuir 2005, 21, 6656–6661. [Google Scholar] [CrossRef]
  167. Yue, B.H.; Huang, C.Y.; Nieh, M.P.; Glinka, C.J.; Katsaras, J. Highly stable phospholipid unilamellar vesicles from spontaneous vesiculation: A DLS and SANS study. J. Phys. Chem. B 2005, 109, 609–616. [Google Scholar] [CrossRef]
  168. Mahabir, S.; Wan, W.; Katsaras, J.; Nieh, M.P. Effects of Charge Density and Thermal History on the Morphologies of Spontaneously Formed Unilamellar Vesicles. J. Phys. Chem. B 2010, 114, 5729–5735. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  169. Frewein, M.P.K.; Doktorova, M.; Heberle, F.A.; Scott, H.L.; Semeraro, E.F.; Porcar, L.; Pabst, G. Structure and Interdigitation of Chain-Asymmetric Phosphatidylcholines and Milk Sphingomyelin in the Fluid Phase. Symmetry 2021, 13, 1441. [Google Scholar] [CrossRef] [PubMed]
  170. Luchini, A.; Cavasso, D.; Radulescu, A.; D’Errico, G.; Paduano, L.; Vitiello, G. Structural Organization of Cardiolipin-Containing Vesicles as Models of the Bacterial Cytoplasmic Membrane. Langmuir 2021, 37, 8508–8516. [Google Scholar] [CrossRef]
  171. Kiselev, M.A.; Zemlyanaya, E.V.; Ryabova, N.Y.; Hauss, T.; Almasy, L.; Funari, S.S.; Zbytovska, J.; Lombardo, D. Influence of ceramide on the internal structure and hydration of the phospholipid bilayer studied by neutron and X-ray scattering. Appl. Phys. A-Mater. Sci. Process. 2014, 116, 319–325. [Google Scholar] [CrossRef]
  172. Sakuma, Y.; Urakami, N.; Taniguchi, T.; Imai, M. Asymmetric distribution of cone-shaped lipids in a highly curved bilayer revealed by a small angle neutron scattering technique. J. Phys.-Condens. Matter 2011, 23, 284104. [Google Scholar] [CrossRef]
  173. Eicher, B.; Heberle, F.A.; Marquardt, D.; Rechberger, G.N.; Katsaras, J.; Pabst, G. Joint small-angle X-ray and neutron scattering data analysis of asymmetric lipid vesicles. J. Appl. Crystallogr. 2017, 50, 419–429. [Google Scholar] [CrossRef] [Green Version]
  174. Eicher, B.; Marquardt, D.; Heberle, F.A.; Letofsky-Papst, I.; Rechberger, G.N.; Appavou, M.S.; Katsaras, J.; Pabst, G. Intrinsic Curvature-Mediated Transbilayer Coupling in Asymmetric Lipid Vesicles. Biophys. J. 2018, 114, 146–157. [Google Scholar] [CrossRef] [Green Version]
  175. Liu, Y.; Kelley, E.G.; Batchu, K.C.; Porcar, L.; Perez-Salas, U. Creating Asymmetric Phospholipid Vesicles via Exchange With Lipid-Coated Silica Nanoparticles. Langmuir 2020, 36, 8865–8873. [Google Scholar] [CrossRef] [PubMed]
  176. Frewein, M.P.K.; Piller, P.; Semeraro, E.F.; Batchu, K.C.; Heberle, F.A.; Scott, H.L.; Gerelli, Y.; Porcar, L.; Pabst, G. Interdigitation-Induced Order and Disorder in Asymmetric Membranes. J. Membr. Biol. 2022, 255, 407–421. [Google Scholar] [CrossRef] [PubMed]
  177. Hirai, M.; Iwase, H.; Hayakawa, T.; Koizumi, M.; Takahashi, H. Determination of Asymmetric Structure of Ganglioside-DPPC Mixed Vesicle Using SANS, SAXS, and DLS. Biophys. J. 2003, 85, 1600–1610. [Google Scholar] [CrossRef]
  178. Petruzielo, R.S.; Heberle, F.A.; Drazba, P.; Katsaras, J.; Feigenson, G.W. Phase behavior and domain size in sphingomyelin-containing lipid bilayers. Biochim. Biophys. Acta-Biomembr. 2013, 1828, 1302–1313. [Google Scholar] [CrossRef] [PubMed]
  179. Heberle, F.A.; Doktorova, M.; Goh, S.L.; Standaert, R.F.; Katsaras, J.; Feigenson, G.W. Hybrid and Nonhybrid Lipids Exert Common Effects on Membrane Raft Size and Morphology. J. Am. Chem. Soc. 2013, 135, 14932–14935. [Google Scholar] [CrossRef] [PubMed]
  180. Heberle, F.A.; Marquardt, D.; Doktorova, M.; Geier, B.; Standaert, R.F.; Heftberger, P.; Kollmitzer, B.; Nickels, J.D.; Dick, R.A.; Feigenson, G.W.; et al. Subnanometer Structure of an Asymmetric Model Membrane: Interleaflet Coupling Influences Domain Properties. Langmuir 2016, 32, 5195–5200. [Google Scholar] [CrossRef]
  181. DiPasquale, M.; Deering, T.G.; Desai, D.; Sharma, A.K.; Amin, S.; Fox, T.E.; Kester, M.; Katsaras, J.; Marquardt, D.; Heberle, F.A. Influence of ceramide on lipid domain stability studied with small-angle neutron scattering: The role of acyl chain length and unsaturation. Chem. Phys. Lipids 2022, 245, 105205. [Google Scholar] [CrossRef]
  182. Ahmadi, D.; Thompson, K.C.; Garcia Sakai, V.; Schweins, R.; Moulin, M.; Haertlein, M.; Strohmeier, G.A.; Pichler, H.; Forsyth, V.T.; Barlow, D.J.; et al. Nanoscale Structure and Dynamics of Model Membrane Lipid Raft Systems, Studied by Neutron Scattering Methods. Front. Phys. 2022, 10, 864746. [Google Scholar] [CrossRef]
  183. Nickels, J.D.; Cheng, X.; Mostofian, B.; Stanley, C.; Lindner, B.; Heberle, F.A.; Perticaroli, S.; Feirgenson, M.; Egami, T.; Standaert, R.F.; et al. Mechanical Properties of Nanoscopic Lipid Domains. J. Am. Chem. Soc. 2015, 137, 15772–15780. [Google Scholar] [CrossRef]
  184. Dante, S.; Thomas Hauß, T.; Brandt, A.; Dencher, N.A. Membrane Fusogenic Activity of the Alzheimer’s Peptide Aβ(1–42) Demonstrated by Small-Angle Neutron Scattering. J. Mol. Biol. 2008, 376, 393–404. [Google Scholar] [CrossRef]
  185. Anunciado, D.; Rai, D.K.; Qian, S.; Urban, V.; O’Neill, H. Small-angle neutron scattering reveals the assembly of alpha-synuclein in lipid membranes. Biochim. Biophys. Acta-Proteins Proteom. 2015, 1854, 1881–1889. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  186. Martel, A.; Antony, L.; Gerelli, Y.; Porcar, L.; Fluitt, A.; Hoffmann, K.; Kiesel, I.; Vivaudou, M.; Fragneto, G.; de Pablo, J.J. Membrane Permeation versus Amyloidogenicity: A Multitechnique Study of Islet Amyloid Polypeptide Interaction with Model Membranes. J. Am. Chem. Soc. 2017, 139, 137–148. [Google Scholar] [CrossRef] [PubMed]
  187. Rai, D.K.; Sharma, V.K.; Anunciado, D.; O’Neill, H.; Mamontov, E.; Urban, V.; Heller, W.T.; Qian, S. Neutron Scattering Studies of the Interplay of Amyloid beta Peptide(1-40) and An Anionic Lipid 1,2-dimyristoyl-sn-glycero-3-phosphoglycerol. Sci. Rep. 2016, 6, 30983. [Google Scholar] [CrossRef] [PubMed]
  188. Rai, D.K.; Qian, S. Interaction of the Antimicrobial Peptide Aurein 1.2 and Charged Lipid Bilayer. Sci. Rep. 2017, 7, 3719. [Google Scholar] [CrossRef] [Green Version]
  189. Nielsen, J.E.; Bjøørnestad, V.A.; Lund, R. Resolving the structural interactions between antimicrobial peptides and lipid membranes using small-angle scattering methods: The case of indolicidin. Soft Matter 2018, 14, 8750–8763. [Google Scholar] [CrossRef]
  190. Rubinson, K.A.; Pokalsky, C.; Krueger, S.; Prochaska, L.J. Structure Determination of Functional Membrane Proteins using Small-Angle Neutron Scattering (SANS) with Small, Mixed-Lipid Liposomes: Native Beef Heart Mitochondrial Cytochrome c Oxidase Forms Dimers. Protein J. 2012, 32, 27–38. [Google Scholar] [CrossRef]
  191. Satsoura, D.; Kučerka, N.; Shivakumar, S.; Pencer, J.; Griffiths, C.; Leber, B.; Andrews, D.W.; Katsaras, J.; Fradin, C. Interaction of the full-length Bax protein with biomimetic mitochondrial liposomes: A small-angle neutron scattering and fluorescence study. Biochim. Biophys. Acta-Biomembr. 2012, 1818, 384–401. [Google Scholar] [CrossRef]
  192. Doktorova, M.; Heberle, F.A.; Kingston, R.L.; Khelashvili, G.; Cuendet, M.A.; Wen, Y.; Katsaras, J.; Feigenson, G.W.; Vogt, V.M.; Dick, R.A. Cholesterol Promotes Protein Binding by Affecting Membrane Electrostatics and Solvation Properties. Biophys. J. 2017, 113, 2004–2015. [Google Scholar] [CrossRef] [Green Version]
  193. Persistence of Vision Pty. Ltd. The PyMOL Molecular Graphics System, Version 3.7; Persistence of Vision Pty. Ltd.: Williamstown, Australia, 2021. [Google Scholar]
  194. Sharma, V.K.; Mamontov, E.; Tyagi, M.; Qian, S.; Rai, D.K.; Urban, V.S. Dynamical and Phase Behavior of a Phospholipid Membrane Altered by an Antimicrobial Peptide at Low Concentration. J. Phys. Chem. Lett. 2016, 7, 2394–2401. [Google Scholar] [CrossRef]
  195. Santamaria, A.; Batchu, K.C.; Matsarskaia, O.; Prévost, S.F.; Russo, D.; Natali, F.; Seydel, T.; Hoffmann, I.; Laux, V.; Haertlein, M.; et al. Strikingly Different Roles of SARS-CoV-2 Fusion Peptides Uncovered by Neutron Scattering. J. Am. Chem. Soc. 2022, 144, 2968–2979. [Google Scholar] [CrossRef]
  196. Ricci, C.; Maccarini, M.; Falus, P.; Librizzi, F.; Mangione, M.R.; Moran, O.; Ortore, M.G.; Schweins, R.; Vilasi, S.; Carrotta, R. Amyloid β-Peptide Interaction with Membranes: Can Chaperones Change the Fate? J. Phys. Chem. B 2019, 123, 631–638. [Google Scholar] [CrossRef] [PubMed]
  197. Kelley, E.G.; Butler, P.D.; Nagao, M. Collective dynamics in lipid membranes containing transmembrane peptides. Soft Matter 2021, 17, 5671–5681. [Google Scholar] [CrossRef] [PubMed]
  198. Kiselev, M.A.; Wartewig, S.; Janich, M.; Lesieur, P.; Kiselev, A.M.; Ollivon, M.; Neubert, R. Does sucrose influence the properties of DMPC vesicles? Chem. Phys. Lipids 2003, 123, 31–44. [Google Scholar] [CrossRef]
  199. Andersen, H.D.; Wang, C.; Arleth, L.; Peters, G.H.; Westh, P. Reconciliation of opposing views on membrane-sugar interactions. Proc. Natl. Acad. Sci. USA 2011, 108, 1874–1878. [Google Scholar] [CrossRef] [Green Version]
  200. Garvey, C.J.; Lenne, T.; Koster, K.L.; Kent, B.; Bryant, G. Phospholipid Membrane Protection by Sugar Molecules during Dehydration-Insights into Molecular Mechanisms Using Scattering Techniques. Int. J. Mol. Sci. 2013, 14, 8148–8163. [Google Scholar] [CrossRef] [Green Version]
  201. Smith, M.D.; Pingali, S.V.; Elkins, J.G.; Bolmatov, D.; Standaert, R.F.; Nickels, J.D.; Urban, V.S.; Katsaras, J.; Davison, B.H.; Smith, J.C.; et al. Solvent-induced membrane stress in biofuel production: Molecular insights from small-angle scattering and all-atom molecular dynamics simulations. Green Chem. 2020, 22, 8278–8288. [Google Scholar] [CrossRef]
  202. Kiselev, M.A.; Lesieur, P.; Kisselev, A.M.; Grabielle-Madelmond, C.; Ollivon, M. DMSO-induced dehydration of DPPC membranes studied by X-ray diffraction, small-angle neutron scattering, and calorimetry. J. Alloys Compd. 1999, 286, 195–202. [Google Scholar] [CrossRef] [Green Version]
  203. Gorshkova, J.E.; Gordeliy, V.I. Investigation of the interaction of dimethyl sulfoxide with lipid membranes by small-angle neutron scattering. Crystallogr. Rep. 2007, 52, 535–539. [Google Scholar] [CrossRef]
  204. Kiselev, M.A.; Zemlyanaya, E.V. Dimethyl sulfoxide-induced dehydration of the intermembrane space of dipalmitoylphosphatidylcholine multilamellar vesicles: Neutron and synchrotron diffraction study. Crystallogr. Rep. 2017, 62, 763–767. [Google Scholar] [CrossRef]
  205. Terakosolphan, W.; Trick, J.L.; Royall, P.G.; Rogers, S.E.; Lamberti, O.; Lorenz, C.D.; Forbes, B.; Harveylle, R.D. Glycerol Solvates DPPC Headgroups and Localizes in the Interfacial Regions of Model Pulmonary Interfaces Altering Bilayer Structure. Langmuir 2018, 34, 6941–6954. [Google Scholar] [CrossRef] [Green Version]
  206. Ristori, S.; Oberdisse, J.; Grillo, I.; Donati, A.; Spalla, O. Structural Characterization of Cationic Liposomes Loaded with Sugar-Based Carboranes. Biophys. J. 2005, 88, 535–547. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  207. Uhríková, D.; Kučerka, N.; Islamov, A.; Gordeliy, V.; Balgavý, P. Small-angle neutron scattering study of N-Dodecyl-N,N-dimethylamine N-oxide induced solubilization of dioleoylphosphatidylcholine bilayers in liposomes. Gen. Physiol. Biophys. 2001, 20, 183–189. [Google Scholar] [PubMed]
  208. Nicolini, C.; Thiyagarajan, P.; Winter, R. Small-scale composition fluctuations and microdomain formation in lipid raft models as revealed by small-angle neutron scattering. Phys. Chem. Chem. Phys. 2004, 6, 5531–5534. [Google Scholar] [CrossRef]
  209. Mkam Tsengam, I.K.; Omarova, M.; Kelley, E.G.; McCormick, A.; Bothun, G.D.; Raghavan, S.R.; John, V.T. Transformation of Lipid Vesicles into Micelles by Adding Nonionic Surfactants: Elucidating the Structural Pathway and the Intermediate Structures. J. Phys. Chem. B 2022, 126, 2208–2216. [Google Scholar] [CrossRef]
  210. Uhríková, D.; Balgavý, P.; Kučerka, N.; Islamov, A.; Gordeliy, V.; Kuklin, A. Small-angle neutron scattering study of the n-decane effect on the bilayer thickness in extruded unilamellar dioleoylphosphatidylcholine liposomes. Biophys. Chem. 2000, 88, 165–170. [Google Scholar] [CrossRef]
  211. Uhríková, D.; Kučerka, N.; Islamov, A.; Kuklin, A.; Gordeliy, V.; Balgavý, P. Small-angle neutron scattering study of the lipid bilayer thickness in unilamellar dioleoylphosphatidylcholine vesicles prepared by the cholate dilution method: N-decane effect. Biochim. Biophys. Acta-Biomembr. 2003, 1611, 31–34. [Google Scholar] [CrossRef] [Green Version]
  212. Usuda, H.; Hishida, M.; Kelley, E.G.; Yamamura, Y.; Nagao, M.; Saito, K. Interleaflet coupling of n-alkane incorporated bilayers. Phys. Chem. Chem. Phys. 2020, 22, 5418–5426. [Google Scholar] [CrossRef]
  213. Misuraca, L.; Calio, A.; LoRicco, J.G.; Hoffmann, I.; Winter, R.; Deme, B.; Peters, J.; Oger, P.M. Alkanes as Membrane Regulators of the Response of Early Membranes to Extreme Temperatures. Life 2022, 12, 445. [Google Scholar] [CrossRef]
  214. Wassall, S.R.; Leng, X.; Canner, S.W.; Pennington, E.R.; Kinnun, J.J.; Cavazos, A.T.; Dadoo, S.; Johnson, D.; Heberle, F.A.; Katsaras, J.; et al. Docosahexaenoic acid regulates the formation of lipid rafts: A unified view from experiment and simulation. Biochim. Biophys. Acta-Biomembr. 2018, 1860, 1985–1993. [Google Scholar] [CrossRef]
  215. Uhríková, D.; Rapp, G.; Yaradaikin, S.; Gordeliy, V.; Balgavý, P. Influence of local anesthetics on the phosphatidylcholine model membrane: Small-angle synchrotron X-ray diffraction and neutron scattering study. Biophys. Chem. 2004, 109, 361–373. [Google Scholar] [CrossRef]
  216. Yi, Z.; Nagao, M.; Bossev, D.P. Effect of charged lidocaine on static and dynamic properties of model bio-membranes. Biophys. Chem. 2012, 160, 20–27. [Google Scholar] [CrossRef] [PubMed]
  217. Mangiapia, G.; Gvaramia, M.; Kuhrts, L.; Teixeira, J.; Koutsioubas, A.; Soltwedel, O.; Frielinghaus, H. Effect of benzocaine and propranolol on phospholipid-based bilayers. Phys. Chem. Chem. Phys. 2017, 19, 32057–32071. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  218. DiPasquale, M.; Nguyen, M.H.L.; Rickeard, B.W.; Cesca, N.; Tannous, C.; Castillo, S.R.; Katsaras, J.; Kelley, E.G.; Heberle, F.A.; Marquardt, D. The antioxidant vitamin E as a membrane raft modulator: Tocopherols do not abolish lipid domains. Biochim. Biophys. Acta-Biomembr. 2020, 1862, 183189. [Google Scholar] [CrossRef] [PubMed]
  219. Drolle, E.; Kucerka, N.; Hoopes, M.I.; Choi, Y.; Katsaras, J.; Karttunen, M.; Leonenko, Z. Effect of melatonin and cholesterol on the structure of DOPC and DPPC membranes. Biochim. Biophys. Acta-Biomembr. 2013, 1828, 2247–2254. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  220. Bolmatov, D.; McClintic, W.T.; Taylor, G.; Stanley, C.B.; Do, C.; Collier, C.P.; Leonenko, Z.; Lavrentovich, M.O.; Katsaras, J. Deciphering Melatonin-Stabilized Phase Separation in Phospholipid Bilayers. Langmuir 2019, 35, 12236–12245. [Google Scholar] [CrossRef]
  221. Sreij, R.; Dargel, C.; Moleiro, L.H.; Monroy, F.; Hellweg, T. Aescin Incorporation and Nanodomain Formation in DMPC Model Membranes. Langmuir 2017, 33, 12351–12361. [Google Scholar] [CrossRef]
  222. Sreij, R.; Prévost, S.; Dargel, C.; Dattani, R.; Hertle, Y.; Wrede, O.; Hellweet, T. Interaction of the Saponin Aescin with Ibuprofen in DMPC Model Membranes. Mol. Pharm. 2018, 15, 4446–4461. [Google Scholar] [CrossRef]
  223. Sreij, R.; Dargel, C.; Hannappel, Y.; Jestin, J.; Prévost, S.; Dattani, R.; Wrede, O.; Hellweg, T. Temperature dependent self-organization of DMPC membranes promoted by intermediate amounts of the saponin aescin. Biochim. Biophys. Acta-Biomembr. 2019, 1861, 897–906. [Google Scholar] [CrossRef]
  224. Sreij, R.; Dargel, C.; Schweins, R.; Prévost, S.; Dattani, R.; Hellweg, T. Aescin-Cholesterol Complexes in DMPC Model Membranes: A DSC and Temperature-Dependent Scattering Study. Sci. Rep. 2019, 9, 5542. [Google Scholar] [CrossRef] [Green Version]
  225. Dargel, C.; Graebitz-Braeuer, F.; Geisler, R.; Fandrich, P.; Hannappel, Y.; Porcar, L.; Hellweg, T. Stable DOPG/Glycyrrhizin Vesicles with a Wide Range of Mixing Ratios: Structure and Stability as Seen by Scattering Experiments and Cryo-TEM. Molecules 2021, 26, 4959. [Google Scholar] [CrossRef]
  226. Minnelli, C.; Moretti, P.; Laudadio, E.; Gerelli, Y.; Pigozzo, A.; Armeni, T.; Galeazzi, R.; Mariani, P.; Mobbili, G. Tuning curvature and phase behavior of monoolein bilayers by epigallocatechin-3-gallate: Structural insight and cytotoxicity. Colloids Surf. B-Biointerfaces 2022, 209, 112171. [Google Scholar] [CrossRef] [PubMed]
  227. Boggara, M.B.; Krishnamoorti, R. Small-Angle Neutron Scattering Studies of Phospholipid-NSAID Adducts. Langmuir 2010, 26, 5734–5745. [Google Scholar] [CrossRef] [PubMed]
  228. De Mel, J.U.; Gupta, S.; Harmon, S.; Stingaciu, L.; Roth, E.W.; Siebenbuerger, M.; Bleuel, M.; Schneider, G.J. Acetaminophen Interactions with Phospholipid Vesicles Induced Changes in Morphology and Lipid Dynamics. Langmuir 2021, 37, 9560–9570. [Google Scholar] [CrossRef] [PubMed]
  229. Foglia, F.; Drake, A.F.; Terry, A.E.; Rogers, S.E.; Lawrence, M.J.; Barlow, D.J. Small-angle neutron scattering studies of the effects of amphotericin B on phospholipid and phospholipid-sterol membrane structure. Biochim. Biophys. Acta-Biomembr. 2011, 1808, 1574–1580. [Google Scholar] [CrossRef]
  230. Hubbard, A.T.M.; Barker, R.; Rehal, R.; Vandera, K.K.A.; Harvey, R.D.; Coates, A.R.M. Mechanism of Action of a Membrane-Active Quinoline-Based Antimicrobial on Natural and Model Bacterial Membranes. Biochemistry 2017, 56, 1163–1174. [Google Scholar] [CrossRef]
  231. Khadka, N.K.; Cheng, X.; Ho, C.S.; Katsaras, J.; Pan, J. Interactions of the Anticancer Drug Tamoxifen with Lipid Membranes. Biophys. J. 2015, 108, 2492–2501. [Google Scholar] [CrossRef] [Green Version]
  232. Ashok, B.; Arleth, L.; Hjelm, R.P.; Rubinstein, I.; Önyüksel, H. In vitro characterization of PEGylated phospholipid micelles for improved drug solubilization: Effects of PEG chain length and PC incorporation. J. Pharm. Sci. 2004, 93, 2476–2487. [Google Scholar] [CrossRef]
  233. Xiao, Y.; Liu, Q.; Clulow, A.J.; Li, T.; Manohar, M.; Gilbert, E.P.; de Campo, L.; Hawley, A.; Boyd, B.J. PEGylation and surface functionalization of liposomes containing drug nanocrystals for cell-targeted delivery. Colloids Surf. B Biointerfaces 2019, 182, 110362. [Google Scholar] [CrossRef]
  234. Arteta, M.Y.; Kjellman, T.; Bartesaghi, S.; Wallin, S.; Wu, X.; Kvist, A.J.; Dabkowska, A.; Székely, N.; Radulescu, A.; Bergenholtz, J.; et al. Successful reprogramming of cellular protein production through mRNA delivered by functionalized lipid nanoparticles. Proc. Natl. Acad. Sci. USA 2018, 115, E3351–E3360. [Google Scholar] [CrossRef] [Green Version]
  235. Nielsen, J.E.; König, N.; Yang, S.; Skoda, M.W.A.; Maestro, A.; Dong, H.; Cárdenas, M.; Lund, R. Lipid membrane interactions of self-assembling antimicrobial nanofibers: Effect of PEGylation. RSC Adv. 2020, 10, 35329–35340. [Google Scholar] [CrossRef]
  236. Tuyen Dao, T.P.; Fernandes, F.; Er-Rafik, M.; Salva, R.; Schmutz, M.; Brûlet, A.; Prieto, M.; Sandre, O.; Le Meins, J.F. Phase Separation and Nanodomain Formation in Hybrid Polymer/Lipid Vesicles. ACS Macro Lett. 2015, 4, 182–186. [Google Scholar] [CrossRef] [Green Version]
  237. Morandi, M.I.; Kluzek, M.; Wolff, J.; Schroder, A.; Thalmann, F.; Marques, C.M. Accumulation of styrene oligomers alters lipid membrane phase order and miscibility. Proc. Natl. Acad. Sci. USA 2021, 118, e2016037118. [Google Scholar] [CrossRef] [PubMed]
  238. Houston, J.E.; Kraft, M.; Mooney, I.; Terry, A.E.; Scherf, U.; Evans, R.C. Charge-Mediated Localization of Conjugated Polythiophenes in Zwitterionic Model Cell Membranes. Langmuir 2016, 32, 8141–8153. [Google Scholar] [CrossRef]
  239. Yu, J.; Mao, J.; Nagao, M.; Bu, W.; Lin, B.; Hong, K.; Jiang, Z.; Liu, Y.; Qian, S.; Tirrell, M.; et al. Structure and dynamics of lipid membranes interacting with antivirulence end-phosphorylated polyethylene glycol block copolymers. Soft Matter 2020, 16, 983–989. [Google Scholar] [CrossRef] [PubMed]
  240. Jiang, T.; Hall, A.; Eres, M.; Hemmatian, Z.; Qiao, B.; Zhou, Y.; Ruan, Z.; Couse, A.D.; Heller, W.T.; Huang, H.; et al. Single-chain heteropolymers transport protons selectively and rapidly. Nature 2020, 577, 216–220. [Google Scholar] [CrossRef] [PubMed]
  241. Richter, A.G.; Dergunov, S.A.; Ganus, B.; Thomas, Z.; Pingali, S.V.; Urban, V.; Liu, Y.; Porcar, L.; Pinkhassik, E. Scattering Studies of Hydrophobic Monomers in Liposomal Bilayers: An Expanding Shell Model of Monomer Distribution. Langmuir 2011, 27, 3792–3797. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  242. Hirai, M.; Onai, T.; Koizumi, M.; Hirai, H.; Kasahara, K.; Suzuki, N.; Yuyama, K.; Inoue, K. Permeability of water through a raft model membrane clarified by time-resolved SANS and SAXS. J. Appl. Crystallogr. 2006, 40, s159–s164. [Google Scholar] [CrossRef] [Green Version]
  243. Mahabir, S.; Small, D.; Li, M.; Wan, W.; Kučerka, N.; Littrell, K.; Katsaras, J.; Nieh, M.P. Growth kinetics of lipid-based nanodiscs to unilamellar vesicles—A time-resolved small angle neutron scattering (SANS) study. Biochim. Biophys. Acta (BBA)-Biomembr. 2013, 1828, 1025–1035. [Google Scholar] [CrossRef]
  244. Nakano, M.; Nakao, H.; Yoshida, S.; Fukuda, M.; Imai, M.; Ikeda, K. Energetic and Structural Insights into Phospholipid Transfer from Membranes with Different Curvatures by Time-Resolved Neutron Scattering. J. Phys. Chem. Lett. 2022, 13, 6024–6030. [Google Scholar] [CrossRef]
  245. Nguyen, M.H.L.; DiPasquale, M.; Rickeard, B.W.; Stanley, C.B.; Kelley, E.G.; Marquardt, D. Methanol Accelerates DMPC Flip-Flop and Transfer: A SANS Study on Lipid Dynamics. Biophys. J. 2019, 116, 755–759. [Google Scholar] [CrossRef] [Green Version]
  246. Garg, S.; Porcar, L.; Woodka, A.; Butler, P.; Perez-Salas, U. Noninvasive Neutron Scattering Measurements Reveal Slower Cholesterol Transport in Model Lipid Membranes. Biophys. J. 2011, 101, 370–377. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  247. Breidigan, J.M.; Krzyzanowski, N.; Liu, Y.; Porcar, L.; Perez-Salas, U. Influence of the membrane environment on cholesterol transfer. J. Lipid Res. 2017, 58, 2255–2263. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  248. Garg, S.; Liu, Y.; Perez-Salas, U.; Porcar, L.; Butler, P.D. Anomalous inter-membrane cholesterol transport in fluid phase phosphoserine vesicles driven by headgroup ordered to disordered entropic transition. Chem. Phys. Lipids 2019, 223, 104779. [Google Scholar] [CrossRef]
  249. Nguyen, M.H.L.; DiPasquale, M.; Rickeard, B.W.; Doktorova, M.; Heberle, F.A.; Scott, H.L.; Barrera, F.N.; Taylor, G.; Collier, C.P.; Stanley, C.B.; et al. Peptide-Induced Lipid Flip-Flop in Asymmetric Liposomes Measured by Small Angle Neutron Scattering. Langmuir 2019, 35, 11735–11744. [Google Scholar] [CrossRef]
  250. Nielsen, J.E.; Lund, R. Molecular Transport and Growth of Lipid Vesicles Exposed to Antimicrobial Peptides. Langmuir 2022, 38, 374–384. [Google Scholar] [CrossRef] [PubMed]
  251. Nielsen, J.E.; Prévost, S.F.; Jenssen, H.; Lund, R. Impact of antimicrobial peptides on E. coli-mimicking lipid model membranes: Correlating structural and dynamic effects using scattering methods. Faraday Discuss. 2021, 232, 203–217. [Google Scholar] [CrossRef]
  252. Masanori, K.; Hiroyuki, N.; Hirokazu, Y.; Hitoshi, E.; Yasushi, I.; Tetsurou, H.; Minoru, N. Control of phospholipid flip-flop by transmembrane peptides. Chem. Phys. 2013, 419, 78–83. [Google Scholar] [CrossRef]
  253. Nguyen, M.H.L.; DiPasquale, M.; Rickeard, B.W.; Yip, C.G.; Greco, K.N.; Kelley, E.G.; Marquardt, D. Time-resolved SANS reveals pore-forming peptides cause rapid lipid reorganization. New J. Chem. 2021, 45, 447–456. [Google Scholar] [CrossRef]
  254. Marx, L.; Frewein, M.P.K.; Semeraro, E.F.; Rechberger, G.N.; Lohner, K.; Porcar, L.; Pabst, G. Antimicrobial peptide activity in asymmetric bacterial membrane mimics. Faraday Discuss. 2021, 232, 435–447. [Google Scholar] [CrossRef]
  255. Nakao, H.; Kimura, Y.; Sakai, A.; Ikeda, K.; Nakano, M. Development of membrane-insertable lipid scrambling peptides: A time-resolved small-angle neutron scattering study. Struct. Dyn. 2021, 8, 024301. [Google Scholar] [CrossRef]
  256. Evans, B.R.; Shah, R. Development of Approaches for Deuterium Incorporation in Plants. Methods Enzymol. 2015, 565, 213–243. [Google Scholar] [CrossRef] [PubMed]
  257. Nagy, G.; Posselt, D.; Kovács, L.; Holm, J.K.; Szabó, M.; Ughy, B.; Rosta, L.; Peters, J.; Timmins, P.; Garab, G. Reversible membrane reorganizations during photosynthesis in vivo: Revealed by small-angle neutron scattering. Biochem. J. 2011, 436, 225–230. [Google Scholar] [CrossRef] [Green Version]
  258. Nagy, G.; Szabó, M.; Ünnep, R.; Káli, G.; Miloslavina, Y.; Lambrev, P.H.; Zsiros, O.; Porcar, L.; Timmins, P.; Rosta, L.; et al. Modulation of the multilamellar membrane organization and of the chiral macrodomains in the diatom Phaeodactylum tricornutum revealed by small-angle neutron scattering and circular dichroism spectroscopy. Photosynth. Res. 2012, 111, 71–79. [Google Scholar] [CrossRef] [PubMed]
  259. Liberton, M.; Page, L.E.; O’Dell, W.B.; O’Neill, H.; Mamontov, E.; Urban, V.S.; Pakrasi, H.B. Organization and Flexibility of Cyanobacterial Thylakoid Membranes Examined by Neutron Scattering. J. Biol. Chem. 2013, 288, 3632–3640. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  260. Liberton, M.; Collins, A.M.; Page, L.E.; O’Dell, W.B.; O’Neill, H.; Urban, V.S.; Timlin, J.A.; Pakrasi, H.B. Probing the consequences of antenna modification in cyanobacteria. Photosynth. Res. 2013, 118, 17–24. [Google Scholar] [CrossRef]
  261. Nagy, G.; Ünnep, R.; Zsiros, O.; Tokutsu, R.; Takizawa, K.; Porcar, L.; Moyet, L.; Petroutsos, D.; Garab, G.; Finazzi, G.; et al. Chloroplast remodeling during state transitions in Chlamydomonas reinhardtii as revealed by noninvasive techniques in vivo. Proc. Natl. Acad. Sci. USA 2014, 111, 5042–5047. [Google Scholar] [CrossRef] [Green Version]
  262. Ünnep, R.; Zsiros, O.; Solymosi, K.; Kovács, L.; Lambrev, P.H.; Tóth, T.; Schweins, R.; Posselt, D.; Székely, N.K.; Rosta, L.; et al. The ultrastructure and flexibility of thylakoid membranes in leaves and isolated chloroplasts as revealed by small-angle neutron scattering. Biochim. Biophys. Acta-Bioenerg. 2014, 1837, 1572–1580. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  263. Stingaciu, L.R.; O’Neill, H.; Liberton, M.; Urban, V.S.; Pakrasi, H.B.; Ohl, M. Revealing the Dynamics of Thylakoid Membranes in Living Cyanobacterial Cells. Sci. Rep. 2016, 6, 19627. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  264. Stingaciu, L.R.; O’Neill, H.M.; Liberton, M.; Pakrasi, H.B.; Urban, V.S. Influence of Chemically Disrupted Photosynthesis on Cyanobacterial Thylakoid Dynamics in Synechocystis sp. PCC 6803. Sci. Rep. 2019, 9, 5711. [Google Scholar] [CrossRef] [Green Version]
  265. Zsiros, O.; Nagy, V.; Párducz, A.; Nagy, G.; Ünnep, R.; El-Ramady, H.; Prokisch, J.; Lisztes-Szabó, Z.; Fári, M.; Csajbók, J.; et al. Effects of selenate and red Se-nanoparticles on the photosynthetic apparatus of Nicotiana tabacum. Photosynth. Res. 2019, 139, 449–460. [Google Scholar] [CrossRef]
  266. Ünnep, R.; Paul, S.; Zsiros, O.; Kovács, L.; Székely, N.K.; Steinbach, G.; Appavou, M.S.; Porcar, L.; Holzwarth, A.R.; Garab, G.; et al. Thylakoid membrane reorganizations revealed by small-angle neutron scattering of Monstera deliciosaleaves associated with non-photochemical quenching. Open Biol. 2020, 10, 200144. [Google Scholar] [CrossRef] [PubMed]
  267. Kanna, S.D.; Domonkos, I.; Kóbori, T.O.; Dergez, A.; Böde, K.; Nagyapáti, S.; Zsiros, O.; Ünnep, R.; Nagy, G.; Garab, G.; et al. Salt Stress Induces Paramylon Accumulation and Fine-Tuning of the Macro-Organization of Thylakoid Membranes in Euglena gracilis Cells. Front. Plant Sci. 2021, 12, 725699. [Google Scholar] [CrossRef]
  268. Murugova, T.N.; Solodovnikova, I.M.; Yurkov, V.I.; Gordeliy, V.I.; Kuklin, A.I.; Ivankov, O.I.; Kovalev, Y.S.; Popov, V.I.; Teplova, V.V.; Yaguzhinsky, L.S. Potentials of Small-angle Neutron Scattering for Studies of the Structure of “Live” Mitochondria. Neutron News 2011, 22, 11–14. [Google Scholar] [CrossRef]
  269. Semeraro, E.F.; Devos, J.M.; Porcar, L.; Forsyth, V.T.; Narayanan, T. In vivo analysis of the Escherichia coli ultrastructure by small-angle scattering. IUCrJ 2017, 4, 751–757. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  270. Semeraro, E.F.; Marx, L.; Mandl, J.; Frewein, M.P.K.; Scott, H.L.; Prévost, S.; Bergler, H.; Lohner, K.; Pabst, G. Evolution of the analytical scattering model of live Escherichia coli. J. Appl. Crystallogr. 2021, 54, 473–485. [Google Scholar] [CrossRef] [PubMed]
  271. Nickels, J.D.; Chatterjee, S.; Stanley, C.B.; Qian, S.; Cheng, X.; Myles, D.A.A.; Standaert, R.F.; Elkins, J.G.; Katsaras, J. The in vivo structure of biological membranes and evidence for lipid domains. PLoS Biol. 2017, 15, e2002214. [Google Scholar] [CrossRef]
Figure 1. Chemical structures of (A) DOPC, (B) DSPE, (C) POPG, (D) SOPS, (E) DMPA, (F) cardiolipin, (G) sphingomyelin, and (H) cholesterol.
Figure 1. Chemical structures of (A) DOPC, (B) DSPE, (C) POPG, (D) SOPS, (E) DMPA, (F) cardiolipin, (G) sphingomyelin, and (H) cholesterol.
Biomolecules 12 01591 g001
Figure 2. Definition of the scattering vector q , which is shown in red.
Figure 2. Definition of the scattering vector q , which is shown in red.
Biomolecules 12 01591 g002
Figure 3. (A) SANS data from 7:3 2 H 54 -DMPC:DMPE at 55 C in 50% 2 H 2 O (black squares), 70% 2 H 2 O (red circles), and 90% 2 H 2 O (blue triangles). The data were collected using the EQ-SANS instrument [88] and have not been previously published. (B) SANS (black squares) and SAXS data (red circles) of 7:3 DMPC:DMPG vesicles in 2 H 2 O. The SANS data were taken from a previously published study [89], while the SAXS data were collected from Sector 12-ID-B of the Advanced Photon Source of Argonne National Laboratory and have not been previously published.
Figure 3. (A) SANS data from 7:3 2 H 54 -DMPC:DMPE at 55 C in 50% 2 H 2 O (black squares), 70% 2 H 2 O (red circles), and 90% 2 H 2 O (blue triangles). The data were collected using the EQ-SANS instrument [88] and have not been previously published. (B) SANS (black squares) and SAXS data (red circles) of 7:3 DMPC:DMPG vesicles in 2 H 2 O. The SANS data were taken from a previously published study [89], while the SAXS data were collected from Sector 12-ID-B of the Advanced Photon Source of Argonne National Laboratory and have not been previously published.
Biomolecules 12 01591 g003
Figure 4. Example of fitting data using Equation (5). The SANS data from Figure 3B are the black squares, while the results of the fit using Equation (5) are shown as the solid red line. T v is 36.4 Å. The low Q portion of the data was not fit because the oscillations resulting from the form factor of the entire vesicle are visible.
Figure 4. Example of fitting data using Equation (5). The SANS data from Figure 3B are the black squares, while the results of the fit using Equation (5) are shown as the solid red line. T v is 36.4 Å. The low Q portion of the data was not fit because the oscillations resulting from the form factor of the entire vesicle are visible.
Biomolecules 12 01591 g004
Figure 5. Schematic of a simple four-shell model showing how the four layers correspond to regions of the bilayer. The layer thicknesses are represented by T i . The lipid was rendered using PyMOL 2.5.2 [97].
Figure 5. Schematic of a simple four-shell model showing how the four layers correspond to regions of the bilayer. The layer thicknesses are represented by T i . The lipid was rendered using PyMOL 2.5.2 [97].
Biomolecules 12 01591 g005
Figure 6. Schematic of parameters used in the self-consistent slab model showing how the lipid layer thickness might be divided into HG and acyl chain regions (left), and the area per lipid used to determine D H C and the number of waters associated with the lipid HG (right). The lipid was rendered using PyMOL 2.5.2 [97].
Figure 6. Schematic of parameters used in the self-consistent slab model showing how the lipid layer thickness might be divided into HG and acyl chain regions (left), and the area per lipid used to determine D H C and the number of waters associated with the lipid HG (right). The lipid was rendered using PyMOL 2.5.2 [97].
Biomolecules 12 01591 g006
Figure 7. Temperature-dependent domain formation in multicomponent phospholipid bilayers. (A,B) represent a schematic of a transition from a laterally uniform vesicle to a state in which there are lateral domains with different SLDs. (CE) show low temperatures (open squares) and high temperatures (solid squares) that reveal the growth of the domains. The figure is Figure 1 of reference [182] by Ahmadi et al. which was published under the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0/, accessed on 22 September 2022). The authors of reference [182] hold the copyright for the figure.
Figure 7. Temperature-dependent domain formation in multicomponent phospholipid bilayers. (A,B) represent a schematic of a transition from a laterally uniform vesicle to a state in which there are lateral domains with different SLDs. (CE) show low temperatures (open squares) and high temperatures (solid squares) that reveal the growth of the domains. The figure is Figure 1 of reference [182] by Ahmadi et al. which was published under the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0/, accessed on 22 September 2022). The authors of reference [182] hold the copyright for the figure.
Biomolecules 12 01591 g007
Figure 8. Schematic of the use of SANS with contrast variation and selective deuteration to determine how charged lipids respond to the peptide melittin [101]. DMPC:DMPS vesicles prepared at a molar ratio of 7:3 with and without the peptide melittin are shown with a wedge removed so that the inner leaflet of the bilayer can be seen. A zoomed-in view is also provided. (A) When no peptide is present, the outer leaflet of the bilayer is enhanced in charged lipids (the light structures) relative to neutral lipids (red and black structures). (B) When melittin is added at a peptide-to-lipid ratio of 1/200, further enhancement of the charged lipid content of the outer leaflet of the bilayer can be seen. The images in the figure used the results of the study in reference [101] and were created using the software developed for the table of contents graphic of the reference and the POV-Ray software version 3.7 [193]. The figure has not been published previously.
Figure 8. Schematic of the use of SANS with contrast variation and selective deuteration to determine how charged lipids respond to the peptide melittin [101]. DMPC:DMPS vesicles prepared at a molar ratio of 7:3 with and without the peptide melittin are shown with a wedge removed so that the inner leaflet of the bilayer can be seen. A zoomed-in view is also provided. (A) When no peptide is present, the outer leaflet of the bilayer is enhanced in charged lipids (the light structures) relative to neutral lipids (red and black structures). (B) When melittin is added at a peptide-to-lipid ratio of 1/200, further enhancement of the charged lipid content of the outer leaflet of the bilayer can be seen. The images in the figure used the results of the study in reference [101] and were created using the software developed for the table of contents graphic of the reference and the POV-Ray software version 3.7 [193]. The figure has not been published previously.
Biomolecules 12 01591 g008
Figure 9. Impact of the amyloid β peptide on the structure and dynamics of DMPG membranes. (A) SANS data showing how the structure of the bilayer changes when peptide is absent from (orange) or present in (cyan) DMPG vesicles. (B) QENS data at two different q-values of DMPG vesicles with the amyloid β peptide. The figure is an adaptation of Figures 1a and 2a of reference [187] by Rai et al., which was published under the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0/, accessed on 22 September 2022). The authors of reference [187] hold the copyright for the figure.
Figure 9. Impact of the amyloid β peptide on the structure and dynamics of DMPG membranes. (A) SANS data showing how the structure of the bilayer changes when peptide is absent from (orange) or present in (cyan) DMPG vesicles. (B) QENS data at two different q-values of DMPG vesicles with the amyloid β peptide. The figure is an adaptation of Figures 1a and 2a of reference [187] by Rai et al., which was published under the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0/, accessed on 22 September 2022). The authors of reference [187] hold the copyright for the figure.
Biomolecules 12 01591 g009
Figure 10. Peptide-enhanced exchange of lipids in vesicles [255]. (A) The integrated intensity as a function of time showing that two peptides engineered to scramble the lipids in the bilayer (4SQ and 4EQ) cause greater lipid exchange than when no peptide is present. (B) The decay in scattering contrast with time for the same system. The image is Figure 3 of reference [255] by Nakao et al., which was published under the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0/, accessed on 22 September 2022). The authors of reference [255] hold the copyright for the figure.
Figure 10. Peptide-enhanced exchange of lipids in vesicles [255]. (A) The integrated intensity as a function of time showing that two peptides engineered to scramble the lipids in the bilayer (4SQ and 4EQ) cause greater lipid exchange than when no peptide is present. (B) The decay in scattering contrast with time for the same system. The image is Figure 3 of reference [255] by Nakao et al., which was published under the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0/, accessed on 22 September 2022). The authors of reference [255] hold the copyright for the figure.
Biomolecules 12 01591 g010
Figure 11. SANS data from a dark-adapted plant leaf showing the response of the thylakoid membranes to light cycling [266]. The image is Figure 1 of reference [266] by Ünnep et al., which was published under the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0/, accessed on 22 September 2022). The authors of reference [266] hold the copyright for the figure.
Figure 11. SANS data from a dark-adapted plant leaf showing the response of the thylakoid membranes to light cycling [266]. The image is Figure 1 of reference [266] by Ünnep et al., which was published under the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0/, accessed on 22 September 2022). The authors of reference [266] hold the copyright for the figure.
Biomolecules 12 01591 g011
Figure 12. SANS with contrast matching applied to a living cell membrane [271]. (a) By feeding the cell deuterium-labeled fatty acids, it was possible to highlight the acyl core of the lipid bilayer. (b) The resulting SANS data could be fit for the thickness of the acyl chain region of the cell membrane. The image is Figure 3 of reference [271] by Nickels et al., which was published under the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0/, accessed on 22 September 2022). The authors of reference [271] hold the copyright for the figure.
Figure 12. SANS with contrast matching applied to a living cell membrane [271]. (a) By feeding the cell deuterium-labeled fatty acids, it was possible to highlight the acyl core of the lipid bilayer. (b) The resulting SANS data could be fit for the thickness of the acyl chain region of the cell membrane. The image is Figure 3 of reference [271] by Nickels et al., which was published under the Creative Commons Attribution License (https://creativecommons.org/licenses/by/4.0/, accessed on 22 September 2022). The authors of reference [271] hold the copyright for the figure.
Biomolecules 12 01591 g012
Table 1. Coherent scattering lengths ( b c o h ) of isotopes of interest for biological membranes [52]. If no specific isotope is noted, the value provided is for the natural abundance of stable isotopes.
Table 1. Coherent scattering lengths ( b c o h ) of isotopes of interest for biological membranes [52]. If no specific isotope is noted, the value provided is for the natural abundance of stable isotopes.
Isotope b coh (fm)
1 H−3.74
2 H6.67
C6.65
N9.36
O5.81
P5.13
Table 2. Example neutron SLDs encountered in SANS studies of biomembranes.
Table 2. Example neutron SLDs encountered in SANS studies of biomembranes.
Chemical or Group ρ ( 10 6  Å 2 )
water
H 2 O−0.560
2 H 2 O6.384
phospholipid HGs
PC HG1.877
2 H 18 -PC HG7.736
PE HG2.483
PA HG2.494
PG HG (in H 2 O)2.382
PG HG (in 2 H 2 O)3.076
PS HG (in H 2 O)3.466
PS HG (in 2 H 2 O)3.893
acyl chains
   in phospholipids
lauroyl (12:0)−0.386
2 H 23 -lauroyl (12:0)6.755
myristoyl (14:0)−0.375
2 H 27 -myristoyl (14:0)6.837
palmitoyl (16:0)−0.386
2 H 31 -palmitoyl (16:0)6.557
oleoyl (18:1c)−0.386
other examples
cholesterol (in H 2 O)0.216
cholesterol (in 2 H 2 O)0.386
melittin (in H 2 O)1.444
melittin (in 2 H 2 O)2.882
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Heller, W.T. Small-Angle Neutron Scattering for Studying Lipid Bilayer Membranes. Biomolecules 2022, 12, 1591. https://doi.org/10.3390/biom12111591

AMA Style

Heller WT. Small-Angle Neutron Scattering for Studying Lipid Bilayer Membranes. Biomolecules. 2022; 12(11):1591. https://doi.org/10.3390/biom12111591

Chicago/Turabian Style

Heller, William T. 2022. "Small-Angle Neutron Scattering for Studying Lipid Bilayer Membranes" Biomolecules 12, no. 11: 1591. https://doi.org/10.3390/biom12111591

Note that from the first issue of 2016, this journal uses article numbers instead of page numbers. See further details here.

Article Metrics

Back to TopTop