Lipid bilayer mechanics

Last updated

Lipid bilayer mechanics is the study of the physical material properties of lipid bilayers, classifying bilayer behavior with stress and strain rather than biochemical interactions. Local point deformations such as membrane protein interactions are typically modelled with the complex theory of biological liquid crystals but the mechanical properties of a homogeneous bilayer are often characterized in terms of only three mechanical elastic moduli: the area expansion modulus Ka, a bending modulus Kb and an edge energy . For fluid bilayers the shear modulus is by definition zero, as the free rearrangement of molecules within plane means that the structure will not support shear stresses. These mechanical properties affect several membrane-mediated biological processes. In particular, the values of Ka and Kb affect the ability of proteins and small molecules to insert into the bilayer. [1] [2] Bilayer mechanical properties have also been shown to alter the function of mechanically activated ion channels. [3]

Contents

Area expansion modulus

Since lipid bilayers are essentially a two dimensional structure, Ka is typically defined only within the plane. Intuitively, one might expect that this modulus would vary linearly with bilayer thickness as it would for a thin plate of isotropic material. In fact this is not the case and Ka is only weakly dependent on bilayer thickness. The reason for this is that the lipids in a fluid bilayer rearrange easily so, unlike a bulk material where the resistance to expansion comes from intermolecular bonds, the resistance to expansion in a bilayer is a result of the extra hydrophobic area exposed to water upon pulling the lipids apart. [4] Based on this understanding, a good first approximation of Ka for a monolayer is 2γ, where gamma is the surface tension of the water-lipid interface. Typically gamma is in the range of 20-50mJ/m2. [5] To calculate Ka for a bilayer it is necessary to multiply the monolayer value by two, since a bilayer is composed of two monolayer leaflets. Based on this calculation, the estimate of Ka for a lipid bilayer should be 80-200 mN/m (note: N/m is equivalent to J/m2). It is not surprising given this understanding of the forces involved that studies have shown that Ka varies strongly with solution conditions [6] but only weakly with tail length and unsaturation. [7]

The compression modulus is difficult to measure experimentally because of the thin, fragile nature of bilayers and the consequently low forces involved. One method utilized has been to study how vesicles swell in response to osmotic stress. This method is, however, indirect and measurements can be perturbed by polydispersity in vesicle size. [6] A more direct method of measuring Ka is the pipette aspiration method, in which a single giant unilamellar vesicle (GUV) is held and stretched with a micropipette. [8] More recently, atomic force microscopy (AFM) has been used to probe the mechanical properties of suspended bilayer membranes, [9] but this method is still under development.

One concern with all of these methods is that, since the bilayer is such a flexible structure, there exist considerable thermal fluctuations in the membrane at many length scales down to sub-microscopic. Thus, forces initially applied to an unstressed membrane are not actually changing the lipid packing but are rather “smoothing out” these undulations, resulting in erroneous values for mechanical properties. [7] This can be a significant source of error. Without the thermal correction typical values for Ka are 100-150 mN/m and with the thermal correction this would change to 220-270 mN/m.

Bending modulus

Bending modulus is defined as the energy required to deform a membrane from its natural curvature to some other curvature. For an ideal bilayer the intrinsic curvature is zero, so this expression is somewhat simplified. The bending modulus, compression modulus and bilayer thickness are related by such that if two of these parameters are known the other can be calculated. This relationship derives from the fact that to bend the inner face must be compressed and the outer face must be stretched. [4] The thicker the membrane, the more each face must deform to accommodate a given curvature (see bending moment). Many of the values for Ka in literature have actually been calculated from experimentally measured values of Kb and t. This relation holds only for small deformations, but this is generally a good approximation as most lipid bilayers can support only a few percent strain before rupturing. [10]

Cross section view of the structures that can be formed by phospholipids in aqueous solutions Phospholipids aqueous solution structures.svg
Cross section view of the structures that can be formed by phospholipids in aqueous solutions

Curvature

Only certain classes of lipids can form bilayers. Two factors primarily govern whether a lipid will form a bilayer or not: solubility and shape. For a self assembled structure such as a bilayer to form, the lipid should have a low solubility in water, which can also be described as a low critical micelle concentration (CMC). [5] Above the CMC, molecules will aggregate and form larger structures such as bilayers, micelles or inverted micelles.

The primary factor governing which structure a given lipid forms is its shape (i.e.- its intrinsic curvature). [4] Intrinsic curvature is defined by the ratio of the diameter of the head group to that of the tail group. For two-tailed PC lipids, this ratio is nearly one so the intrinsic curvature is nearly zero. Other headgroups such as PS and PE are smaller and the resulting diacyl (two-tailed) lipids thus have a negative intrinsic curvature. Lysolipids tend to have positive spontaneous curvature because they have one rather than two alkyl chains in the tail region. If a particular lipid has too large a deviation from zero intrinsic curvature it will not form a bilayer. [11]

Edge energy

Schematic showing two possible conformations of the lipids at the edge of a pore. In the top image the lipids have not rearranged, so the pore wall is hydrophobic. In the bottom image some of the lipid heads have bent over, so the pore wall is hydrophilic. Pore schematic.svg
Schematic showing two possible conformations of the lipids at the edge of a pore. In the top image the lipids have not rearranged, so the pore wall is hydrophobic. In the bottom image some of the lipid heads have bent over, so the pore wall is hydrophilic.

Edge energy is the energy per unit length of a free edge contacting water. This can be thought of as the work needed to create a hole in the bilayer of unit length L. The origin of this energy is the fact that creating such an interface exposes some of the lipid tails to water, which is unfavorable. is also an important parameter in biological phenomena as it regulates the self-healing properties of the bilayer following electroporation or mechanical perforation of the cell membrane. [8] Unfortunately, this property is both difficult to measure experimentally and to calculate. One of the major difficulties in calculation is that the structural properties of this edge are not known. The simplest model would be no change in bilayer orientation, such that the full length of the tail is exposed. This is a high energy conformation and, to stabilize this edge, it is likely that some of the lipids rearrange their head groups to point out in a curved boundary.[ citation needed ] The extent to which this occurs is currently unknown and there is some evidence that both hydrophobic (tails straight) and hydrophilic (heads curved around) pores can coexist. [12]

Finite element modeling

FE Modeling is a powerful tool for testing the mechanical deformation and equilibrium configuration of lipid membranes. [13] In this context membranes are treated under the thin-shell theory where the bending behavior of the membrane is described by the Helfrich bending model which considers the bilayer as being a very thin object and interprets it as a two-dimensional surface. This consideration imply that Kirchhoff-Love plate theory can be applied to lipid bilayers to determine their stress-deformation behavior. Furthermore, in the FE approach a bilayer surface is subdivided into discrete elements, each described by the above 2D mechanics. [14] [15]

Under these considerations the weak form virtual work for the entire system is described as the sum of the contribution of all the discrete elements work components. [15]

For each discrete element the virtual work is determined by the force vector and the displacement vector each for an applied stress and a bending momentum

The FE force vectors due to the applied bilayer stress are given as

Here is the displacement state function at point , and the tangent vector to the bilayer surface at point

The above individual element vectors for the internal force and the internal work can be expressed in a global assembly to obtain a discretized weak form as follows:

In the above equation is the deformation in each discrete element while is the Lagrange multiplier associated with area-incompressibility. The discretized weak form is satisfied when and . The resulting nonlinear equations are solved using Newton ‘s method. This allows the predictions of equilibrium shapes that lipid membranes adopt under different stimuli. [14] [16] [17]

Most analysis are done for lipid membranes with uniform properties (isotropic), while this is partly true for simple membranes containing a single or few lipid species, this description can only approximate the mechanical response of more complex lipid bilayers which can contain several domains of segregated lipids having distinct material properties or intermembrane proteins as in the case of cellular membranes. Other complex cases requiring surface flow analysis, pH, and temperature dependency would require a more discrete model of a bilayer such as MD simulations. [16] FE methods can predict the equilibrium conformations of a lipid bilayer in response to external forces as shown in the following cases.

Tethering

In this scenario a point in the bilayer surface is pulled with a force normal to the surface plane, this leads to the elongation of a thin tether of bilayer material. The rest of the bilayer surface is subject to a tension S reflecting the pulling force of the continuous bilayer. In such this case a finer mesh is applied near the pulling force area to have a more accurate prediction of the bilayer deformation. [16] Tethering is an important mechanical event for cellular lipid bilayers, by this action membranes are able to mediate docking into substrates or components of the cytoskeleton.

Budding

Lipid bilayer budding is a commonplace phenomenon in living cells and relates to the transport of metabolites in the form of vesicles. During this process, a lipid bilayer is subject to internal hydrostatic stresses, in combination with strain restrictions along a bilayer surface, this can lead to elongation of areas of the lipid bilayer by elastic shear or viscous shear. This eventually leads to a deformation of a typical spherical bilayer into different budding shapes. Such shapes are not restricted to be symmetrical along their axes, but they can have different degrees of asymmetry. In FE analysis this results in budding equilibrium shapes like, elongate plates, tubular buds and symmetric budding. [16]

Related Research Articles

Continuum mechanics is a branch of mechanics that deals with the deformation of and transmission of forces through materials modeled as a continuous medium rather than as discrete particles. The French mathematician Augustin-Louis Cauchy was the first to formulate such models in the 19th century.

<span class="mw-page-title-main">Lipid bilayer</span> Membrane of two layers of lipid molecules

The lipid bilayer is a thin polar membrane made of two layers of lipid molecules. These membranes are flat sheets that form a continuous barrier around all cells. The cell membranes of almost all organisms and many viruses are made of a lipid bilayer, as are the nuclear membrane surrounding the cell nucleus, and membranes of the membrane-bound organelles in the cell. The lipid bilayer is the barrier that keeps ions, proteins and other molecules where they are needed and prevents them from diffusing into areas where they should not be. Lipid bilayers are ideally suited to this role, even though they are only a few nanometers in width, because they are impermeable to most water-soluble (hydrophilic) molecules. Bilayers are particularly impermeable to ions, which allows cells to regulate salt concentrations and pH by transporting ions across their membranes using proteins called ion pumps.

<span class="mw-page-title-main">Hooke's law</span> Physical law: force needed to deform a spring scales linearly with distance

In physics, Hooke's law is an empirical law which states that the force needed to extend or compress a spring by some distance scales linearly with respect to that distance—that is, Fs = kx, where k is a constant factor characteristic of the spring, and x is small compared to the total possible deformation of the spring. The law is named after 17th-century British physicist Robert Hooke. He first stated the law in 1676 as a Latin anagram. He published the solution of his anagram in 1678 as: ut tensio, sic vis. Hooke states in the 1678 work that he was aware of the law since 1660.

<span class="mw-page-title-main">Poisson's ratio</span> Measure of material deformation perpendicular to loading

In materials science and solid mechanics, Poisson's ratio (nu) is a measure of the Poisson effect, the deformation of a material in directions perpendicular to the specific direction of loading. The value of Poisson's ratio is the negative of the ratio of transverse strain to axial strain. For small values of these changes, is the amount of transversal elongation divided by the amount of axial compression. Most materials have Poisson's ratio values ranging between 0.0 and 0.5. For soft materials, such as rubber, where the bulk modulus is much higher than the shear modulus, Poisson's ratio is near 0.5. For open-cell polymer foams, Poisson's ratio is near zero, since the cells tend to collapse in compression. Many typical solids have Poisson's ratios in the range of 0.2–0.3. The ratio is named after the French mathematician and physicist Siméon Poisson.

<span class="mw-page-title-main">Liposome</span> Composite structures made of phospholipids and may contain small amounts of other molecules

A liposome is a small artificial vesicle, spherical in shape, having at least one lipid bilayer. Due to their hydrophobicity and/or hydrophilicity, biocompatibility, particle size and many other properties, liposomes can be used as drug delivery vehicles for administration of pharmaceutical drugs and nutrients, such as lipid nanoparticles in mRNA vaccines, and DNA vaccines. Liposomes can be prepared by disrupting biological membranes.

In materials science and continuum mechanics, viscoelasticity is the property of materials that exhibit both viscous and elastic characteristics when undergoing deformation. Viscous materials, like water, resist shear flow and strain linearly with time when a stress is applied. Elastic materials strain when stretched and immediately return to their original state once the stress is removed.

<span class="mw-page-title-main">Dislocation</span> Linear crystallographic defect or irregularity

In materials science, a dislocation or Taylor's dislocation is a linear crystallographic defect or irregularity within a crystal structure that contains an abrupt change in the arrangement of atoms. The movement of dislocations allow atoms to slide over each other at low stress levels and is known as glide or slip. The crystalline order is restored on either side of a glide dislocation but the atoms on one side have moved by one position. The crystalline order is not fully restored with a partial dislocation. A dislocation defines the boundary between slipped and unslipped regions of material and as a result, must either form a complete loop, intersect other dislocations or defects, or extend to the edges of the crystal. A dislocation can be characterised by the distance and direction of movement it causes to atoms which is defined by the Burgers vector. Plastic deformation of a material occurs by the creation and movement of many dislocations. The number and arrangement of dislocations influences many of the properties of materials.

<span class="mw-page-title-main">Buckling</span> Sudden change in shape of a structural component under load

In structural engineering, buckling is the sudden change in shape (deformation) of a structural component under load, such as the bowing of a column under compression or the wrinkling of a plate under shear. If a structure is subjected to a gradually increasing load, when the load reaches a critical level, a member may suddenly change shape and the structure and component is said to have buckled. Euler's critical load and Johnson's parabolic formula are used to determine the buckling stress of a column.

<span class="mw-page-title-main">Bending</span> Strain caused by an external load

In applied mechanics, bending characterizes the behavior of a slender structural element subjected to an external load applied perpendicularly to a longitudinal axis of the element.

<span class="mw-page-title-main">Euler–Bernoulli beam theory</span> Method for load calculation in construction

Euler–Bernoulli beam theory is a simplification of the linear theory of elasticity which provides a means of calculating the load-carrying and deflection characteristics of beams. It covers the case corresponding to small deflections of a beam that is subjected to lateral loads only. By ignoring the effects of shear deformation and rotatory inertia, it is thus a special case of Timoshenko–Ehrenfest beam theory. It was first enunciated circa 1750, but was not applied on a large scale until the development of the Eiffel Tower and the Ferris wheel in the late 19th century. Following these successful demonstrations, it quickly became a cornerstone of engineering and an enabler of the Second Industrial Revolution.

<span class="mw-page-title-main">Charge density</span> Electric charge per unit length, area or volume

In electromagnetism, charge density is the amount of electric charge per unit length, surface area, or volume. Volume charge density is the quantity of charge per unit volume, measured in the SI system in coulombs per cubic meter (C⋅m−3), at any point in a volume. Surface charge density (σ) is the quantity of charge per unit area, measured in coulombs per square meter (C⋅m−2), at any point on a surface charge distribution on a two dimensional surface. Linear charge density (λ) is the quantity of charge per unit length, measured in coulombs per meter (C⋅m−1), at any point on a line charge distribution. Charge density can be either positive or negative, since electric charge can be either positive or negative.

The J-integral represents a way to calculate the strain energy release rate, or work (energy) per unit fracture surface area, in a material. The theoretical concept of J-integral was developed in 1967 by G. P. Cherepanov and independently in 1968 by James R. Rice, who showed that an energetic contour path integral was independent of the path around a crack.

A cell membrane defines a boundary between a cell and its environment. The primary constituent of a membrane is a phospholipid bilayer that forms in a water-based environment due to the hydrophilic nature of the lipid head and the hydrophobic nature of the two tails. In addition there are other lipids and proteins in the membrane, the latter typically in the form of isolated rafts.

<span class="mw-page-title-main">Contact mechanics</span> Study of the deformation of solids that touch each other

Contact mechanics is the study of the deformation of solids that touch each other at one or more points. A central distinction in contact mechanics is between stresses acting perpendicular to the contacting bodies' surfaces and frictional stresses acting tangentially between the surfaces. Normal contact mechanics or frictionless contact mechanics focuses on normal stresses caused by applied normal forces and by the adhesion present on surfaces in close contact, even if they are clean and dry. Frictional contact mechanics emphasizes the effect of friction forces.

<span class="mw-page-title-main">Mechanical properties of carbon nanotubes</span>

The mechanical properties of carbon nanotubes reveal them as one of the strongest materials in nature. Carbon nanotubes (CNTs) are long hollow cylinders of graphene. Although graphene sheets have 2D symmetry, carbon nanotubes by geometry have different properties in axial and radial directions. It has been shown that CNTs are very strong in the axial direction. Young's modulus on the order of 270 - 950 GPa and tensile strength of 11 - 63 GPa were obtained.

A model lipid bilayer is any bilayer assembled in vitro, as opposed to the bilayer of natural cell membranes or covering various sub-cellular structures like the nucleus. They are used to study the fundamental properties of biological membranes in a simplified and well-controlled environment, and increasingly in bottom-up synthetic biology for the construction of artificial cells. A model bilayer can be made with either synthetic or natural lipids. The simplest model systems contain only a single pure synthetic lipid. More physiologically relevant model bilayers can be made with mixtures of several synthetic or natural lipids.

Membrane curvature is the geometrical measure or characterization of the curvature of membranes. The membranes can be naturally occurring or man-made (synthetic). An example of naturally occurring membrane is the lipid bilayer of cells, also known as cellular membranes. Synthetic membranes can be obtained by preparing aqueous solutions of certain lipids. The lipids will then "aggregate" and form various phases and structures. According to the conditions and the chemical structures of the lipid, different phases will be observed. For instance, the lipid POPC tends to form lamellar vesicles in solution, whereas smaller lipids, such as detergents, will form micelles if the CMC was reached. There are five commonly proposed mechanisms by which membrane curvature is created, maintained, or controlled: lipid composition, shaped transmembrane proteins, protein motif insertion/BAR domains, protein scaffolding, and cytoskeleton scaffolding.

The bidomain model is a mathematical model to define the electrical activity of the heart. It consists in a continuum (volume-average) approach in which the cardiac microstructure is defined in terms of muscle fibers grouped in sheets, creating a complex three-dimensional structure with anisotropical properties. Then, to define the electrical activity, two interpenetrating domains are considered, which are the intracellular and extracellular domains, representing respectively the space inside the cells and the region between them.

<span class="mw-page-title-main">Scallop theorem</span>

In physics, the scallop theorem states that a swimmer that performs a reciprocal motion cannot achieve net displacement in a low-Reynolds number Newtonian fluid environment, i.e. a fluid that is highly viscous. Such a swimmer deforms its body into a particular shape through a sequence of motions and then reverts to the original shape by going through the sequence in reverse. It does not matter how fast or slow the swimmer executes the sequence. At low Reynolds number, time or inertia does not come into play, and the swimming motion is purely determined by the sequence of shapes that the swimmer assumes.

Flickering analysis of cellular or membranous structures is a widespread technique for measuring the bending modulus and other properties from the power spectrum of thermal fluctuations.

References

  1. Garcia ML (July 2004). "Gate expectations". Nature. 430 (6996): 153–5. doi:10.1038/430153a. PMID   15241399. S2CID   4427370.
  2. McIntosh TJ, Simon SA (2006). "Roles of bilayer material properties in function and distribution of membrane proteins". Annual Review of Biophysics and Biomolecular Structure. 35: 177–98. doi:10.1146/annurev.biophys.35.040405.102022. PMID   16689633.
  3. Suchyna TM, Tape SE, Koeppe RE, Andersen OS, Sachs F, Gottlieb PA (July 2004). "Bilayer-dependent inhibition of mechanosensitive channels by neuroactive peptide enantiomers". Nature. 430 (6996): 235–40. Bibcode:2004Natur.430..235S. doi:10.1038/nature02743. PMID   15241420. S2CID   4401688.
  4. 1 2 3 Boal DH (2002). Mechanics of the Cell. Cambridge, UK: Cambridge University Press. ISBN   978-0-521-79681-1.
  5. 1 2 Israelachvili JN (2011). Intermolecular and surface forces (3rd ed.). Burlington, MA: Academic Press. ISBN   978-0-08-092363-5.
  6. 1 2 Rutkowski CA, Williams LM, Haines TH, Cummins HZ (June 1991). "The elasticity of synthetic phospholipid vesicles obtained by photon correlation spectroscopy". Biochemistry. 30 (23): 5688–96. doi:10.1021/bi00237a008. PMID   2043611.
  7. 1 2 Rawicz W, Olbrich KC, McIntosh T, Needham D, Evans E (July 2000). "Effect of chain length and unsaturation on elasticity of lipid bilayers". Biophysical Journal. 79 (1): 328–39. Bibcode:2000BpJ....79..328R. doi:10.1016/S0006-3495(00)76295-3. PMC   1300937 . PMID   10866959.
  8. 1 2 Evans E, Heinrich V, Ludwig F, Rawicz W (October 2003). "Dynamic tension spectroscopy and strength of biomembranes". Biophysical Journal. 85 (4): 2342–50. Bibcode:2003BpJ....85.2342E. doi:10.1016/S0006-3495(03)74658-X. PMC   1303459 . PMID   14507698.
  9. Steltenkamp S, Müller MM, Deserno M, Hennesthal C, Steinem C, Janshoff A (July 2006). "Mechanical properties of pore-spanning lipid bilayers probed by atomic force microscopy". Biophysical Journal. 91 (1): 217–26. Bibcode:2006BpJ....91..217S. doi:10.1529/biophysj.106.081398. PMC   1479081 . PMID   16617084.
  10. Hallett FR, Marsh J, Nickel BG, Wood JM (February 1993). "Mechanical properties of vesicles. II. A model for osmotic swelling and lysis". Biophysical Journal. 64 (2): 435–42. Bibcode:1993BpJ....64..435H. doi:10.1016/S0006-3495(93)81384-5. PMC   1262346 . PMID   8457669.
  11. Zeidi M, Kim CI (August 2018). "The Effects of Intra-membrane Viscosity on Lipid Membrane Morphology: Complete Analytical Solution". Scientific Reports. 8 (1): 12845. Bibcode:2018NatSR...812845Z. doi: 10.1038/s41598-018-31251-6 . PMC   6110749 . PMID   30150612.
  12. Weaver JC, Chizmadzhev YA (December 1996). "Theory of electroporation: a review". Bioelectrochemistry and Bioenergetics. 41 (2): 135–60. doi:10.1016/S0302-4598(96)05062-3.
  13. Kim J (February 2020). "Probing nanomechanical responses of cell membranes". Scientific Reports. 10 (1): 2301. Bibcode:2020NatSR..10.2301K. doi:10.1038/s41598-020-59030-2. PMC   7010710 . PMID   32041981.
  14. 1 2 Steigmann DJ (2018). The role of mechanics in the study of lipid bilayers. Cham. ISBN   978-3-319-56348-0. OCLC   988293321.{{cite book}}: CS1 maint: location missing publisher (link)
  15. 1 2 Sauer RA (2017-05-25). "On the Computational Modeling of Lipid Bilayers Using Thin-Shell Theory". The Role of Mechanics in the Study of Lipid Bilayers. CISM International Centre for Mechanical Sciences. Vol. 577. Cham: Springer International Publishing. pp. 221–286. doi:10.1007/978-3-319-56348-0_5. ISBN   978-3-319-56347-3.
  16. 1 2 3 4 Sauer RA, Duong TX, Mandadapu KK, Steigmann DJ (February 2017). "A stabilized finite element formulation for liquid shells and its application to lipid bilayers". Journal of Computational Physics. 330: 436–466. arXiv: 1601.03907 . Bibcode:2017JCoPh.330..436S. doi:10.1016/j.jcp.2016.11.004. S2CID   357683.
  17. Bidhendi AJ, Geitmann A (January 2018). "Finite Element Modeling of Shape Changes in Plant Cells". Plant Physiology. 176 (1): 41–56. doi:10.1104/pp.17.01684. PMC   5761827 . PMID   29229695.