You are on page 1of 5

pubs.acs.

org/biochemistry Perspective

The Origin of Lipid Rafts


Steven L. Regen*

Cite This: Biochemistry 2020, 59, 4617−4621 Read Online

ACCESS Metrics & More Article Recommendations

ABSTRACT: The time-averaged lateral organization of the lipids and


See https://pubs.acs.org/sharingguidelines for options on how to legitimately share published articles.

proteins that make up mammalian cell membranes continues to be the


subject of intense interest and debate. Since the introduction of the fluid
mosaic model almost 50 years ago, the “lipid raft hypothesis” has emerged as
a popular concept that has captured the imagination of a large segment of
the biomembrane community. In particular, the notion that lipid rafts play a
pivotal role in cellular processes such as signal transduction and membrane
Downloaded via UENF on March 31, 2022 at 13:47:54 (UTC).

protein trafficking is now favored by many investigators. Despite the


attractiveness of lipid rafts, their composition, size, lifetime, biological
function, and even the very existence remain controversial. The central tenet
that underlies this hypothesis is that cholesterol and high-melting lipids have
favorable interactions (i.e., they pull together), which lead to transient
domains. Recent nearest-neighbor recognition (NNR) studies have
expanded the lipid raft hypothesis to include the influence that low-melting
lipids have on the organization of lipid membranes. Specifically, it has been found that mimics of cholesterol and high-melting lipids
are repelled (i.e., pushed away) by low-melting lipids in fluid bilayers. The picture that has emerged from our NNR studies is that
lipid mixing is governed by a balance of these “push and pull” forces, which maximizes the number of hydrocarbon contacts and
attractive van der Waals interactions within the membrane. The power of the NNR methodology is that it allows one to probe these
push/pull interaction energies that are measured in tens of calories per mole.

■ THE LIPID RAFT HYPOTHESIS


Despite a wealth of information that is currently available on the
composition and dynamic properties of mammalian cell
membranes, the time-averaged lateral organization of the lipids
and proteins that make up these natural enclosures remains
controversial.1−4 A concept that has captured the imagination of
many researchers in this area is that high-melting lipids (i.e.,
lipids having a gel to liquid-crystalline phase transition
temperature, Tm, greater than 37 °C) associate with cholesterol
to form transient domains termed “lipid rafts”.5,6 This Figure 1. Stylized illustration of a raft “phase” coexisting with a nonraft
association appears to be the result of attractive van der Waals region. (Reprinted from ref 9.)
interactions between these two molecules.7 It has also been
posited that lipid rafts play a pivotal role in cellular processes
involving signal transduction and membrane protein traffick- studies of analogous bilayers via fluorescence lifetime, differ-
ing.8 A stylized illustration of a hypothetical lipid raft region ential scanning calorimetry,2H NMR, and Forster resonance
coexisting with a nonraft region is shown in Figure 1.9 energy transfer measurements have yielded additional support


for cholesterol favoring association with high-melting lip-
CHOLESTEROL’S AFFINITY TOWARD ids.12−15 Related experiments that we have carried out ourselves
HIGH-MELTING LIPIDS
A large body of evidence now exists that shows that cholesterol Received: October 21, 2020
favors association with high-melting lipids over low-melting Revised: November 13, 2020
lipids (i.e., lipids having Tm values less than 37 °C). This Published: November 23, 2020
evidence can be traced back to early monolayer studies that were
carried out at the air/water interface where discrete cholesterol−
phospholipid complexes were hypothesized.10,11 Subsequent

© 2020 American Chemical Society https://dx.doi.org/10.1021/acs.biochem.0c00851


4617 Biochemistry 2020, 59, 4617−4621
Biochemistry pubs.acs.org/biochemistry Perspective

support a “hydrophobic contact mechanism” as a basis for this homodimers, AA and BB, by a factor 2. When values of K deviate
affinity.7,9 Specifically, our results support a model in which the from 4, this is a measure of what we have termed nearest-
flexible, saturated acyl chains of high-melting lipids establish a neighbor recognition (NNR). Thus, when K is greater than 4,
large number of hydrocarbon contacts with the flat and rigid unlike-nearest-neighbors (i.e., heterodimers) are favored. When
cholesterol molecule, thereby producing strong attractive van K is less than 4, like-nearest-neighbors (i.e., homodimers) are
der Waals interactions.7 In contrast, low-melting phospholipids, favored. The net free energy of the interaction between A and B,
which bear one or more “permanent kinks” (i.e., cis-double ωAB, is then given by ωAB = −1/2RTln(K/4).9 Those values of
bonds), have a weaker association with cholesterol because they ωAB that are greater than 0 cal per mol of phospholipid are
are unable to create as many hydrocarbon contacts.7 considered to be repulsive, which means that unlike lipids prefer

■ EARLY STUDIES INDICATING THE EXISTENCE OF


REPULSIVE FORCES BETWEEN CHOLESTEROL AND
to separate from each other. Values of ωAB that are less than 0 cal
per mol of phospholipid are considered to reflect attractive
interactions between unlike nearest-neighbors.
LOW-MELTING LIPIDS
Using a fluorescence-quenching strategy, it has been shown that
low-melting lipids enhance the stability of liquid-ordered
■ EXCHANGEABLE LIPID MIMICS
Chart 1 shows the high-melting DPPC molecule (Tm = 41.5 °C)
bilayers derived from cholesterol and high-melting lipids such and cholesterol, along with their exchangeable mimics Phos (Tm
as 1,2-dipalmitoyl-sn-glycerol-3-phosphocholine (DPPC).16
Although such experiments do not provide quantitative insight Chart 1. Structures of DPPC, Cholesterol, and Exchangeable
into lipid−lipid interactions, they imply that low-melting lipids Mimics
repel cholesterol and/or DPPC.
In a separate investigation that used FRET measurements and
Monte Carlo simulations to study the interactions among
porcine brain sphingomyelin (SM), cholesterol (Chol), and 1-
palmitoyl-2-oleoyl-sn-glycerol-3-phosphocholine (POPC), it
was concluded that SM/Chol interactions are attractive while
both SM/POPC and Chol/POPC interactions are weakly
repulsive.17

■ NEAREST-NEIGHBOR RECOGNITION
MEASUREMENTS
In an effort to obtain a quantitative assessment of lipid−lipid
interactions in fluid bilayers, we devised a chemical method that
is based on the use of exchangeable, disulfide-based lipid
dimers.9 The essence of this method is illustrated in Figure 2.

Figure 2. Illustration of the nearest-neighbor recognition (NNR) = 41.9 °C) and Chol, respectively.18,19 Chart 2 shows three low-
method. Here, the solid rectangles represent an exchangeable
melting exchangeable lipids, c1-Phos, c2-Phos, and c3-Phos,
cholesterol mimic, and the large filled circles with two wavy lines
represent an exchangeable phospholipid. which we have investigated in bilayers derived from 1-palmitoyl-
2-dihydrosterculoyl-sn-glycero-3-phosphocholine (PDSPC).20
Experimental evidence that supports the claim that these
Thus, a heterodimer, AB, is incorporated in a host membrane exchangeable lipids are excellent mimics of cholesterol and the
that is derived from (i) low-melting phospholipids and is corresponding phosphocholines has been obtained from
maintained in the “liquid-disordered” state or (ii) high-melting differential scanning calorimetry, surface-pressure area isotherm,
phospholipids plus cholesterol that is maintained in the “liquid- fluorescence, and nearest-neighbor recognition measure-
ordered” state; i.e., a state that is thought to mimic lipid rafts. ments.9,21
Partial reduction with dithiothreitol (DTT) then releases As discussed elsewhere, cis-double bonds undergo cis/trans
monomers that undergo exchange with residual dimers via the isomerization under NNR reaction conditions.22 This isomer-
thiolate-disulfide interchange. The dimer distribution is then ization appears to be due to the generation of small amounts of
monitored by high-performance liquid chromatography until thiyl radicals and their reversible addition to the double bonds of
chemical equilibrium has been reached. This equilibrium is phospholipids.23−25 To circumvent this complexity, we have
defined by AA + BB ⇌ 2 AB. When an equilibrium constant (K) used cis-cyclopropyl moieties as a means of introducing
for the exchange is found to equal 4, this indicates that A and B permanent “kinks” in the acyl chains to create low-melting
are mixing ideally; i.e., the monomers are randomly distributed, exchangeable lipids. For example, PDSPC has a low Tm value of
and the heterodimer, AB, is statistically favored over each of the −10 °C, which is similar to the Tm value of POPC of −3 °C.26
4618 https://dx.doi.org/10.1021/acs.biochem.0c00851
Biochemistry 2020, 59, 4617−4621
Biochemistry pubs.acs.org/biochemistry Perspective

Chart 2. Structures of Low-Melting Exchangeable attractive and repulsive forces in fluid bilayers made from
Phospholipids PDSPC (Figure 3).

■ THE ORIGIN OF LIPID RAFT FORMATION


The picture that has emerged from our NNR studies is that lipid
mixing in fluid bilayers is governed by “push and pull” forces.
More specifically, high-melting lipids favor binding to flat, rigid
cholesterol molecules by using their flexible, saturated acyl
chains to make a high number of hydrocarbon contacts with
them.7 For low-melting lipids, the number of hydrocarbon
contacts that can be made with this sterol is significantly reduced
due to permanent kinks in their acyl chains. Despite this
reduction in hydrocarbon contact, it is sufficient for cholesterol
to exert its long-studied condensing effect, as evidenced by
monolayer measurements.27 Thus, in bilayers containing
cholesterol, high-melting lipids, and low-melting lipids, the
number of hydrocarbon contacts and attractive van der Waals
interactions are maximized by having the low-melting lipids
push away cholesterol and also the high-melting lipids, thereby
allowing them to pull together to form less fluid domains, i.e.,


lipid rafts (Figure 4). At the same time, this increased
QUANTIFYING ATTRACTIVE AND REPULSIVE
FORCES
To assess the interactions between Phos and Chol in
membranes that mimic lipid rafts, we used host membranes
made from DPPC and cholesterol (60:40, mol/mol). At 45 °C,
an attractive force was found that corresponded to ωAB = −260
cal/mol of phospholipid.19 In sharp contrast, related NNR
measurements that were carried out in bilayers made from
PDSPC revealed net repulsive interactions between Chol and
the low-melting lipids, c1-Phos, c2-Phos, and c3-Phos (Figure
3).20 In addition, the magnitude of these repulsive interactions

Figure 4. Illustration summarizing the “push and pull” forces among


cholesterol, low-melting lipids, and high-melting lipids that drive the
formation of lipid rafts.

hydrocarbon contact minimizes unfavorable exposure of the


hydrophobic portion of the lipids toward water. In other words,
both van der Waals interactions and the hydrophobic effect
contribute to the formation of lipid rafts.

■ THE HIDDEN ROLE THAT POLYUNSATURATED


LIPIDS ARE LIKELY TO PLAY IN FORMING LIPID
Figure 3. Net free energies of the interaction between Chol and the RAFTS
low-melting lipids, c1-Phos, c2-Phos, and c3-Phos, in host membranes
made from PDSPC at 45 °C. Also shown is the net free energy of the Although substantial attention has focused on favorable
interaction between Chol and Phos in host membranes made from interactions between high-melting lipids and cholesterol in
PDSPC at 45 °C.20 forming lipid rafts over the past 20 years, relatively little
attention has been paid to the role that low-melting lipids may
was found to increase as the number of permanent kinks play. Given the abundance of polyunsaturated lipids in
increased. It is noteworthy that the magnitude of the repulsive
forces between Chol and these low-melting lipids was virtually mammalian cell membranes and the fact that repulsive
identical with those of Phos interacting with these same low- interactions increase with increasing numbers of permanent
melting lipids (Figure 3). Since the net free energy of interaction kinks, it appears likely that the repulsive interactions that
between Chol and Phos in bilayers of PDSPC is essentially 0 cal/ cholesterol and high-melting lipids experience with polyunsatu-
mol, these findings indicate that Chol and Phos have similar rated lipids are a major driving force for lipid raft formation.28−33
4619 https://dx.doi.org/10.1021/acs.biochem.0c00851
Biochemistry 2020, 59, 4617−4621
Biochemistry

■ ■
pubs.acs.org/biochemistry Perspective

PUSH AND PULL FORCES AS A MEASURE OF LIPID AUTHOR INFORMATION


MISCIBILITY Corresponding Author
In this Perspective, I have referred to values of ωAB that are Steven L. Regen − Department of Chemistry, Lehigh University,
greater than 0 cal per mol of phospholipid as being repulsive Bethlehem, Pennsylvania 18015, United States; orcid.org/
(representing a “push”) and values of ωAB that are less than 0 cal 0000-0001-6192-7916; Email: slr0@lehigh.edu
per mol of phospholipid as being attractive (representing a
“pull”). An alternative way of thinking about these push and pull Complete contact information is available at:
forces is to consider them as a measure of miscibility where like- https://pubs.acs.org/10.1021/acs.biochem.0c00851
attracts-like. Thus, a value of ωAB that is strongly positive reflects
two lipids that are dissimilar and do not want to become nearest- Notes
neighbors. In other words, they are immiscible at the molecular The author declares no competing financial interest.
and at the supramolecular level. In contrast, a value of ωAB that is
strongly negative reflects two different lipids that want to
become nearest-neighbors and are miscible. Figure 5 shows how
■ ACKNOWLEDGMENTS
I am most grateful to my co-workers who are listed in our
publications that are cited in this Perspective. I am also grateful
to the National Institutes of Health (PHS GM56149) and the
National Science Foundation (CHE-1145500) for their
financial support, which has made our research in this area
and this Perspective possible.

■ REFERENCES
(1) Singer, S., and Nicolson, G. (1972) The fluid mosaic model of the
structure of cell membranes. Science 175, 720−731.
(2) Engelman, D. M. (2005) Membranes are more mosaic than fluid.
Figure 5. Snapshots of Monte Carlo simulations of mixtures of Chol Nature 438, 578−580.
with (A) c1-Phos and (B) c3-Phos. The mixtures contain 40% Chol (3) Levental, I., Levental, K. R., and Heberle, F. A. (2020) Lipid rafts:
(red) and 60% of c1-Phos (black) or c3-Phos (black). The total number controversies resolved, mysteries remain. Trends Cell Biol. 30 (5), 341−
of lipids in each snapshot is 10,000. (Reprinted from ref 34). 353.
(4) Levental, I. (2020) Lipid rafts come of age. Nat. Rev. Mol. Cell Biol.
21, 420.
(5) Simons, K., and Ikonen, E. (1997) Functional rafts in cell
two different repulsive forces can influence lipid mixing on a membranes. Nature 387 (6633), 569−572.
(6) Ahmed, S. N., Brown, D. A., and London, E. (1997) On the origin
macroscopic scale. Here, snapshots of Monte Carlo simulations of sphingolipid/cholesterol-rich detergent-insoluble cell membranes:
are shown for 40:60 (mol/mol) mixtures of Chol/c1-Phos and physiological concentrations of cholesterol and sphingolipid induce
Chol/c3-Phos. These simulations were based, directly, on our formation of a detergent-insoluble, liquid-ordered phase in model
nearest-neighbor recognition measurements. Each snapshot membranes. Biochemistry 36, 10944−10953.
encompasses a total of 10,000 lipids. In the case of Chol and c1- (7) Daly, T. A., Wang, M., and Regen, S. L. (2011) The origin of
cholesterol’s condensing effect. Langmuir 27, 2159−2161.
Phos, where ωAB = +160 cal/mol, the two lipids are found to mix (8) Lingwood, D., and Simons, K. (2010) Lipid rafts as a membrane-
very well. However, a comparison of this snapshot with a similar organizing principle. Science 327, 46−50.
Monte Carlo snapshot that was based on a value of ωAB = 0 cal/ (9) Krause, M. R., and Regen, S. L. (2014) The structural role of
mol indicates that this mixing is not ideal (not shown).34 In the cholesterol in cell membranes: from condensed bilayers to lipid rafts.
Acc. Chem. Res. 47, 3512−3521.
case of Chol and c3-Phos, where ωAB = +430 cal/mol, the mixing (10) Radhakrishnan, A., and McConnell, H. M. (1999) Cholesterol-
of the two lipids is clearly very poor.34 It is noteworthy, in this phospholipid complexes in membranes. J. Am. Chem. Soc. 121, 486−
regard, that small-angle neutron scattering experiments have 487.
revealed that lipid domain size can increase with the extent of (11) Radhakrishnan, A., and McConnell, H. M. (1999) Condensed
complexes of cholesterol and phospholipids. Biophys. J. 77, 1507−1517.
acyl chain unsaturation.35 (12) Radhakrishnan, A., and McConnell, H. M. (2005) Condensed

■ CONCLUSIONS
Although the composition, size, lifetime, and biological function
complexes in vesicles containing cholesterol and phospholipids. Proc.
Natl. Acad. Sci. U. S. A. 102, 12662−12666.
(13) Engberg, O., Hautala, V., Yasuda, T., Dehio, H., Murata, M.,
Slotte, J. P., and Nyholm, T. K.M. (2016) The affinity of cholesterol for
of lipid rafts remain to be defined, the nearest-neighbor different phospholipids affects lateral segregation in bilayers. Biophys. J.
recognition measurements that I have described in this 111, 546−556.
Perspective provide a basis for their likely existence. Specifically, (14) Nyholm, T. K. M., Engberg, O., Hautala, V., Tsuchikawa, H., Lin,
K.-L., Murata, M., and Slotte, J. P. (2019) Impact of acyl chain
these measurements have revealed the presence of push and pull mismatch on the formation and properties of sphingomyelin-
forces that can maximize attractive van der Waals forces within a cholesterol domains. Biophys. J. 117, 1577−1588.
membrane. Based on the extreme sensitivity of the nearest- (15) Engberg, O., Lin, K.-L., Hautala, V., Slotte, J. P., and Nyholm, T.
neighbor recognition method, which allows for interaction K. M. (2020) Sphingomyelin acyl chains influence the formation of
sphingomyelin- and cholesterol-encriched domains. Biophys. J. 119,
energies to be measured in tens of calories per mole, I believe 913−923.
that, with careful molecular design, it could also provide unique (16) Bakht, O., Pathak, P., and London, E. (2007) Effect of the
insight into lipid-membrane protein interactions.36 structure of lipids favoring disordered domain formation on the stability

4620 https://dx.doi.org/10.1021/acs.biochem.0c00851
Biochemistry 2020, 59, 4617−4621
Biochemistry pubs.acs.org/biochemistry Perspective

of cholesterol-containing ordered domains (lipid rafts): identification (35) Heberle, F. A., Petruzielo, R. S., Pan, J., Drazba, P., Kuceerka, N.,
of multiple raft-stabilization mechanisms. Biophys. J. 93, 4307−4318. Standaert, R. F., Feigenson, G. W., and Katsaras, J. (2013) Bilayer
(17) Frazier, M. L., Wright, J. R., Pokorny, A., and Almeida, P. F. F. thickness mismatch controls domain size in model membranes. J. Am.
(2007) Investigation of domain formation in sphingomyelin/ Chem. Soc. 135, 6853−6859.
cholesterol/popc mixtures by fluorescence resonance energy transfer (36) Huster, D. (2014) Solid-state NMR spectroscopy to study
and monte carlo simulations. Biophys. J. 92, 2422−2433. protein-lipid interactions. Biochim. Biophys. Acta, Mol. Cell Biol. Lipids
(18) Krisovitch, S. M., and Regen, S. L. (1992) Nearest-neighbor 1841, 1146−1160.
recognition in phospholipid membranes: a molecular-level approach to
the study of membrane suprastructure. J. Am. Chem. Soc. 114, 9828−
9835.
(19) Turkyilmaz, S., Almeida, P. F., and Regen, S. L. (2011) Effects of
isoflurane, halothane and chloroform on the interactions and lateral
organization of lipids in the liquid-ordered phase. Langmuir 27, 14380−
14385.
(20) Wang, C., Yu, Y., and Regen, S. L. (2017) Lipid raft formation:
key role of polyunsaturated phospholipids. Angew. Chem., Int. Ed. 56,
1639−1642.
(21) Mukai, M., and Regen, S. L. (2015) Exchangeable mimics of dppc
and dppg exhibiting similar nearest-neighbor interactions in fluid
bilayers. Langmuir 31, 12674−12678.
(22) Dewa, T., Vigmond, S. J., and Regen, S. L. (1996) Lateral
heterogeneity in fluid bilayers composed of saturated and unsaturated
phospholipids. J. Am. Chem. Soc. 118, 3435−3440.
(23) Chatgilialoglu, C., Ferreri, C., Ballestri, M., Mulazzani, Q. G., and
Landi, L. (2000) cis-trans Isomerization of monounsaturated fatty acid
residues in phospholipids by thiyl radicals. J. Am. Chem. Soc. 122, 4593−
4601.
(24) Ferreri, C., Costantino, C., Perrotta, L., Landi, L., Mulazzani, Q.
G., and Chatgilialoglu, C. (2001) cis-trans Isomerization of poly-
unsaturated fatty acid residues in phospholipids catalyzed by thiyl
radicals. J. Am. Chem. Soc. 123, 4459−4468.
(25) Ferreri, C., Samadi, A., Sassatelli, F., Landi, L., and Chatgilialoglu,
C. (2004) Regioselective cis-trans isomerization of arachidonic double
bonds by thiyl radicals: the influence of phospholipid supramolecular
organization. J. Am. Chem. Soc. 126, 1063−1072.
(26) Koynova, R., and Caffrey, M. (1998) Phases and phase
transitions of the phosphatidylcholines. Biochim. Biophys. Acta, Rev.
Biomembr. 1376, 91−145.
(27) Jurak, M. (2013) Thermodynamic aspects of cholesterol effect
on properties of phospholipid monolayers: langmuir and langmuir-
blodgett monolayer study. J. Phys. Chem. B 117, 3496−3501.
(28) Kucerka, N., Marquardt, D., Harroun, T. A., Nieh, M.-P., Wassall,
S. R., de Jong, D. H., Schafer, L. V., Marrink, S. J., and Katsaras, J. (2010)
Cholesterol in bilayers with pufa chains: doping with dmpc or popc
results in sterol reorientation and membrane-domain formation.
Biochemistry 49, 7485−7493.
(29) Shaikh, S. R., Kinnun, J. J., Leng, X., Williams, J. A., and Wassall, S.
R. (2015) How polyunsaturated fatty acids modify molecular
organization in membranes: insight from NMR studies of model
systems. Biochim. Biophys. Acta, Biomembr. 1848, 211−219.
(30) Lin, X., Lorent, J. H., Skinkle, A. D., Levental, K. R., Waxham, M.
N., Gorfe, A. A., and Levental, I. (2016) Domain stability in biomimetic
membranes driven by lipid polyunsaturation. J. Phys. Chem. B 120,
11930−11941.
(31) Levental, K. R., Lorent, J. H., Lin, X., Skinkle, A. D., Surma, M. A.,
Stockenbojer, E. A., Gorfe, A. A., and Levental, I. (2016)
Polyunsaturated lipids regulate membrane domain stability by tuning
membrane order. Biophys. J. 110, 1800−1810.
(32) Bennett, W. F. D., Shea, J.-E., and Tieleman, D. P. (2018)
Phospholipid chain interactions with cholesterol drive domain
formation in lipid membranes. Biophys. J. 114, 2595−2605.
(33) Engberg, O., Hautala, V., Yasuda, T., Dehio, H., Murata, M.,
Slotte, P. J., and Nyholm, T. K. M. (2016) (2016) The affinity of
cholesterol for different phospholipids affects lateral segregation in
bilayers. Biophys. J. 111, 546−556.
(34) Wang, C., Almeida, P. F., and Regen, S. L. (2018) Net
interactions that push cholesterol away from unsaturated phospholipids
are driven by enthalpy. Biochemistry 57, 6637−6643.

4621 https://dx.doi.org/10.1021/acs.biochem.0c00851
Biochemistry 2020, 59, 4617−4621

You might also like