insights into sphingolipid miscibility: separate observation of sphingomyelin and ceramide n-acyl...

10
Insights into Sphingolipid Miscibility: Separate Observation of Sphingomyelin and Ceramide N-Acyl Chain Melting Sherry S. W. Leung, Jon V. Busto, Amir Keyvanloo, Fe ´lix M. Gon ˜i, and Jenifer Thewalt †§ * Department of Physics and § Department of Molecular Biology and Biochemistry, Simon Fraser University, Burnaby, British Columbia, Canada; and Unidad de Biofı ´sica (Centro Mixto Spanish Scientific Research Council CSIC, University of the Basque Country), and Departamento de Bioquı ´mica, Universidad del Paı ´s Vasco, Bilbao, Spain ABSTRACT Ceramide produced from sphingomyelin in the plasma membrane is purported to affect signaling through changes in the membrane’s physical properties. Thermal behavior of N-palmitoyl sphingomyelin (PSM) and N-palmitoyl ceram- ide (PCer) mixtures in excess water has been monitored by 2 H NMR spectroscopy and compared to differential scanning calo- rimetry (DSC) data. The alternate use of either perdeuterated or proton-based N-acyl chain PSM and PCer in our 2 H NMR studies has allowed the separate observation of gel-fluid transitions in each lipid in the presence of the other one, and this in turn has provided direct information on the lipids’ miscibility over a wide temperature range. The results provide further evidence of the stabilization of the PSM gel state by PCer. Moreover, overlapping NMR and DSC data reveal that the DSC-signals parallel the melting of the major component (PSM) except at intermediate (20 and 30 mol %) fractions of PCer. In such cases, the DSC endotherm reports on the presumably highly cooperative melting of PCer. Up to at least 50 mol % PCer, PSM and PCer mix ideally in the liquid crystalline phase; in the gel phase, PCer becomes incorporated into PSM:PCer membranes with no evidence of pure solid PCer. INTRODUCTION The sphingolipid signaling pathway (1,2) is critical, among other things, in directing healthy and tumoral cells toward either programmed death or survival (3,4). An early stage of this pathway consists of the hydrolysis of cell membrane sphingomyelin (SM) by a sphingomyelinase, giving rise to water-soluble phosphorylcholine and highly nonpolar ceramide (Cer) (5,6). Cer remains in the membrane and is responsible for downstream steps in the pathway, either binding specific enzymes, or perhaps changing the physical properties of the membrane (7,8). Thus, as a result of sphin- gomyelinase activity, a mixture of SM and Cer occurs, and lateral separation of phospholipid-enriched and Cer- enriched domains is likely within the cell membrane (9,10). Coalescence of initially small Cer-enriched domains may give rise to micron-sized clusters or platforms, contain- ing specific proteins, and exerting specialized functions, i.e., in the cellular response to stress (11,12). The above discoveries at the model and cellular level, together with previous observations of Cer properties (7,13–15), have led to a number of physical studies specif- ically directed to explore domain formation in bilayers (16–19). Studies of SM:Cer mixtures are less frequent (20). Formation of Cer-enriched domains upon treatment of SM-containing vesicles with a sphingomyelinase was described probably for the first time by Holopainen et al. (21). Further studies from Maggio’s laboratory have provided interesting data on the subject of Cer domains generated by sphingomyelinase. Epifluorescence studies of SM monolayers being degraded by sphingomyelinase, added to the aqueous subphase, indicate in real-time that formation of Cer alters surface topography at an early stage, inducing phase separation into condensed (Cer-enriched) and expanded (SM-enriched) domains (22–24). Cer-en- riched domains in egg SM:Cer bilayers, rather than mono- layers, were described by Sot et al. (25) using mainly differential scanning calorimetry (DSC) and fluorescence microscopy. DSC data showed for pure egg SM a rather narrow gel-fluid transition centered at 39 C. Egg Cer, even at low proportions (5 mol %), gave rise to asymmetric endotherms, indicating the formation of high-temperature melting SM:Cer domains. The corresponding phase diagram includes a wide temperature-composition region of coexist- ing domains. These are also observed in giant unilamellar vesicles of the same composition using fluorescence micros- copy (25). Here we characterize in detail the temperature-composi- tion diagram of chemically defined N-palmitoyl SM:N- palmitoyl Cer (PSM:PCer) mixtures, using 2 H NMR and DSC. The use of such pure lipids enables the drawing of more accurate phase boundaries. 2 H NMR has the advantage that experiments may be performed by deuterating either the SM or the Cer N-acyl chains, so that melting of each mole- cule can be separately observed. DSC provides an overall view of the phase transition, and constitutes a link with previous studies (e.g., Sot et al. (25)). 2 H NMR allowed the first observation of Cer-rich domains in bilayers composed of fully deuterated DPPC and bovine brain Cer (26). We earlier used 2 H NMR to study mixtures of PCer and 1-palmitoyl, 2-oleoyl PC and found that gel and liquid-crystalline (L a ) phases coexisted over a wide range Submitted June 4, 2012, and accepted for publication October 3, 2012. *Correspondence: [email protected] Editor: Klaus Gawrisch. Ó 2012 by the Biophysical Society 0006-3495/12/12/2465/10 $2.00 http://dx.doi.org/10.1016/j.bpj.2012.10.041 Biophysical Journal Volume 103 December 2012 2465–2474 2465

Upload: jenifer

Post on 30-Dec-2016

213 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Insights into Sphingolipid Miscibility: Separate Observation of Sphingomyelin and Ceramide N-Acyl Chain Melting

Biophysical Journal Volume 103 December 2012 2465–2474 2465

Insights into Sphingolipid Miscibility: Separate Observation ofSphingomyelin and Ceramide N-Acyl Chain Melting

Sherry S. W. Leung,† Jon V. Busto,‡ Amir Keyvanloo,† Felix M. Goni,‡ and Jenifer Thewalt†§*†Department of Physics and §Department of Molecular Biology and Biochemistry, Simon Fraser University, Burnaby, British Columbia,Canada; and ‡Unidad de Biofısica (Centro Mixto Spanish Scientific Research Council CSIC, University of the Basque Country), andDepartamento de Bioquımica, Universidad del Paıs Vasco, Bilbao, Spain

ABSTRACT Ceramide produced from sphingomyelin in the plasma membrane is purported to affect signaling throughchanges in the membrane’s physical properties. Thermal behavior of N-palmitoyl sphingomyelin (PSM) and N-palmitoyl ceram-ide (PCer) mixtures in excess water has been monitored by 2H NMR spectroscopy and compared to differential scanning calo-rimetry (DSC) data. The alternate use of either perdeuterated or proton-based N-acyl chain PSM and PCer in our 2H NMRstudies has allowed the separate observation of gel-fluid transitions in each lipid in the presence of the other one, and this inturn has provided direct information on the lipids’ miscibility over a wide temperature range. The results provide further evidenceof the stabilization of the PSM gel state by PCer. Moreover, overlapping NMR and DSC data reveal that the DSC-signals parallelthe melting of the major component (PSM) except at intermediate (20 and 30 mol %) fractions of PCer. In such cases, the DSCendotherm reports on the presumably highly cooperative melting of PCer. Up to at least 50 mol % PCer, PSM and PCer mixideally in the liquid crystalline phase; in the gel phase, PCer becomes incorporated into PSM:PCer membranes with no evidenceof pure solid PCer.

INTRODUCTION

The sphingolipid signaling pathway (1,2) is critical, amongother things, in directing healthy and tumoral cells towardeither programmed death or survival (3,4). An early stageof this pathway consists of the hydrolysis of cell membranesphingomyelin (SM) by a sphingomyelinase, giving riseto water-soluble phosphorylcholine and highly nonpolarceramide (Cer) (5,6). Cer remains in the membrane and isresponsible for downstream steps in the pathway, eitherbinding specific enzymes, or perhaps changing the physicalproperties of the membrane (7,8). Thus, as a result of sphin-gomyelinase activity, a mixture of SM and Cer occurs, andlateral separation of phospholipid-enriched and Cer-enriched domains is likely within the cell membrane(9,10). Coalescence of initially small Cer-enriched domainsmay give rise to micron-sized clusters or platforms, contain-ing specific proteins, and exerting specialized functions, i.e.,in the cellular response to stress (11,12).

The above discoveries at the model and cellular level,together with previous observations of Cer properties(7,13–15), have led to a number of physical studies specif-ically directed to explore domain formation in bilayers(16–19). Studies of SM:Cer mixtures are less frequent(20). Formation of Cer-enriched domains upon treatmentof SM-containing vesicles with a sphingomyelinase wasdescribed probably for the first time by Holopainen et al.(21). Further studies from Maggio’s laboratory haveprovided interesting data on the subject of Cer domainsgenerated by sphingomyelinase. Epifluorescence studies of

Submitted June 4, 2012, and accepted for publication October 3, 2012.

*Correspondence: [email protected]

Editor: Klaus Gawrisch.

� 2012 by the Biophysical Society

0006-3495/12/12/2465/10 $2.00

SM monolayers being degraded by sphingomyelinase,added to the aqueous subphase, indicate in real-time thatformation of Cer alters surface topography at an early stage,inducing phase separation into condensed (Cer-enriched)and expanded (SM-enriched) domains (22–24). Cer-en-riched domains in egg SM:Cer bilayers, rather than mono-layers, were described by Sot et al. (25) using mainlydifferential scanning calorimetry (DSC) and fluorescencemicroscopy. DSC data showed for pure egg SM a rathernarrow gel-fluid transition centered at 39�C. Egg Cer,even at low proportions (5 mol %), gave rise to asymmetricendotherms, indicating the formation of high-temperaturemelting SM:Cer domains. The corresponding phase diagramincludes a wide temperature-composition region of coexist-ing domains. These are also observed in giant unilamellarvesicles of the same composition using fluorescence micros-copy (25).

Here we characterize in detail the temperature-composi-tion diagram of chemically defined N-palmitoyl SM:N-palmitoyl Cer (PSM:PCer) mixtures, using 2H NMR andDSC. The use of such pure lipids enables the drawing ofmore accurate phase boundaries. 2H NMR has the advantagethat experiments may be performed by deuterating either theSM or the Cer N-acyl chains, so that melting of each mole-cule can be separately observed. DSC provides an overallview of the phase transition, and constitutes a link withprevious studies (e.g., Sot et al. (25)). 2H NMR allowedthe first observation of Cer-rich domains in bilayerscomposed of fully deuterated DPPC and bovine brain Cer(26). We earlier used 2H NMR to study mixtures of PCerand 1-palmitoyl, 2-oleoyl PC and found that gel andliquid-crystalline (La) phases coexisted over a wide range

http://dx.doi.org/10.1016/j.bpj.2012.10.041

Page 2: Insights into Sphingolipid Miscibility: Separate Observation of Sphingomyelin and Ceramide N-Acyl Chain Melting

2466 Leung et al.

of temperatures and compositions, with domains of differentcomposition and physical state being present at physiolog-ical temperature (27).

A rational approach to complexity is to start at a simplelevel, upon which more complicated conceptual structurescan be built. A more detailed understanding of lipid-lipidinteractions in the plasma membrane would require, in addi-tion to SM and Cer, the presence of at least cholesterol andunsaturated PC. However this quaternary mixture is hardlyamenable to analysis by contemporary techniques; eventhe ternary mixtures of SM:PC:cholesterol (17,28,29) aredifficult to study, and the mixture SM:Cer:cholesterol isknown for its unusual behavior (30–32). Thus, we startwith SM:Cer, a decision supported by the fact that thesetwo lipids give rise to complexes that are stable even inthe presence of PC or cholesterol under certain conditions(31). This investigation of SM:Cer provides insights intothe miscibility of these two lipids. The data will help inthe analysis of mixtures containing SM:Cer plus cholesteroland/or unsaturated PC.

MATERIALS AND METHODS

Materials

Synthetic N-palmitoyl-D-erythro-sphingosylphosphorylcholine (PSM)

and N-[2H31] palmitoyl-D-erythro-sphingosylphosphorylcholine (PSM-

D31) in powder form were obtained from Avanti Polar Lipids (Alabaster,

AL). N-palmitoyl-D-erythro-sphingosine (PCer) and N-[2H31] palmitoyl-

D-erythro-sphingosine (PCer-D31) in powder form were obtained from

Northern Lipids (Vancouver, British Columbia, Canada). Lipid purity was

checked with thin layer chromatography using 70:26:4 chloroform/meth-

anol/water (v/v). Deuterium-depleted water was purchased from Sigma-

Aldrich (St. Louis, MO).

Multilamellar vesicle preparation

PSM multilamellar vesicles (MLVs) containing nominal PCer concentra-

tions of 0, 10, 20, 30, and 40 mol % were made using standard methods.

Briefly, appropriate quantities of PSM and PCer were combined, dissolved

in 80:20 benzene:methanol (v/v), lyophilized until dry, hydrated in excess

deuterium-depleted water (700 mL), and passed through five freeze-thaw-

vortex cycles between �196 and 95�C (pure PCer melts at 90�C (14)).

One of the two lipids is perdeuterated in the acyl chain for 2H NMR. The

actual MLV compositions were mole ratios of 100:0, 90:10, 80:20, 74:26,

and 60:40 PSM-D31:PCer, and 90:10, 80:20, 69:31, and 61:39

PSM:PCer-D31. The uncertainty in concentration was %2% for PSM-

D31:PCer samples and %0.5% for PSM:PCer-D31 samples. Total lipid

mass varied between 11 and 161 mg.

2H NMR

2H NMR was performed using a 7.0 T Oxford Magnet (Oxford Magnet

Technology, Witney, Oxon, UK) with a locally built spectrometer operating

at 46.8 MHz using the quadrupolar echo technique (33), which involved

two 90� pulses of 3.95 ms duration, 90� out of phase, 40 ms apart, and

repeated every 300 ms. Data were collected with quadrature detection

and 8-CYCLOPS phase cycling. At least 10,000 scans were averaged.

Experiments were performed immediately after MLV preparation. MLVs

were allowed to equilibrate at each temperature for at least 20 min before

Biophysical Journal 103(12) 2465–2474

data collection. Experiments proceeded from low to high temperatures

with 1� increments except at high temperatures to minimize water loss. To

check for formation of a slowly relaxing PCer crystalline phase at tempera-

tures below the gel-La transition (27), quadrupolar echo experiments were

also performed using a repetition time of 50 s. No evidence of a solid phase

was observed. Typical data acquisition for each sample took 3–4 weeks.

The first moment (M1), or average spectral width, is defined for our

symmetric spectrum as

M1 ¼Px

u¼�x

jujf ðuÞ

2Px

u¼�x

f ðuÞ; (1)

where u is the frequency, f(u) is the intensity at u, and x ¼ 2p,40 kHz for

fully La spectra and 2p,75 kHz otherwise.

Each deuteron along a perdeuterated lipid chain in an La phase

membrane contributes a Pake doublet to the spectrum, with an associated

quadrupolar splitting DnQ of

DnQ ¼ 3

4

e2qQ

hjSCDj; (2)

where e2qQ/h is the static quadrupolar coupling constant and SCD is

a measure of the angular excursions of the C-D bond with respect to the

lipid long axis. Gel phase spectra, by contrast, are bell-shaped, due to the

lack of rapid axially symmetric motion of the lipids about their long axes.

In a phase coexistence region, 2H NMR spectra are superpositions of the

weighted gel and La end-point spectra. Given two 2H NMR spectra from

such a region (at the same temperature), the end-point spectra can be

obtained by subtracting a fraction K (or K0) of one spectrum from the other.

Because the gel and La phases have different echo decay rates, the K (or K0)values were corrected according to the procedure in Morrow et al. (34). The

end-point concentrations for the gel and La phases are

xG ¼ ð1� xBÞxA � Kð1� xAÞxBð1� xBÞ � Kð1� xAÞ ; (3)

ð1� xAÞxB � K0ð1� xBÞxA

xL ¼ ð1� xAÞ � K0ð1� xBÞ ; (4)

where xA and xB are the concentration of labeled lipid in samples with more

gel and more La content, respectively (35).

RESULTS

2H NMR spectra

In the PSM:PCer system, the individual behavior of the twosphingolipids was observed by perdeuterating the N-palmi-toyl chain of either one in turn. Aqueous MLVs of PSM-D31:PCer and PSM:PCer-D31 were prepared with 10, 20,30, and 40 mol % PCer (or PCer-D31) and 2H NMR datawere collected from 25 to 80�C. The spectra of purePSM-D31 MLVs from 25 to 70�C were also recorded; thepure PSM-D31 membrane exhibits a sharp gel-La phasetransition at 40�C, in close agreement with the literature(28,36).

Fig. 1 shows the spectra of 90:10 PSM-D31:PCer andPSM:PCer-D31 at select temperatures. Visual inspection

Page 3: Insights into Sphingolipid Miscibility: Separate Observation of Sphingomyelin and Ceramide N-Acyl Chain Melting

-40 0 40kHz

-40 0 40kHz

90:10 Spm-D31:Cer 90:10 Spm:Cer-D31

75°C

65°C

57°C

56°C

47°C

41°C

25°C

51°C

39°C

75°C

65°C

57°C

56°C

47°C

41°C

39°C

25°C

51°C

FIGURE 1 2H NMR spectra for 90:10 PSM-D31:PCer and 90:10

PSM:PCer-D31 at various temperatures. (Boxed temperatures) Boundaries

of the phase transition. The minor baseline irregularities at560 kHz in the

56–75�C spectra of 90:10 PSM:PCer-D31 are artifacts from acoustic

ringing.

1.4

1.2

1.0

0.8

0.6M1 (

105 s

-1)

807060504030Temperature (ºC)

0% 10% 20% 30% 40%

1.4

1.2

1.0

0.8

0.6

M1 (

105 s

-1)

807060504030Temperature (ºC)

10% 20% 30% 40%

FIGURE 2 Effect of PCer concentration on the temperature dependence

of M1 for 100:0(C), 90:10(:), 80:20(;), 70:30(A), 60:40(-) PSM-

D31:PCer (solid symbols), and PSM:PCer-D31 (open symbols).

Insights into Sphingolipid Miscibility 2467

of 2H NMR spectra gives direct information about the stateof the labeled lipid. Gel spectra are broad and rather feature-less, extending to 563 kHz, as is the case at 25�C for bothsamples. In contrast, a number of distinct peaks are presentin a La spectrum, such as those at 75�C. In comparison topure PSM-D31, these mixed sphingolipids undergo a broadtransition. Because experiments progress from low to hightemperatures, we define the lowest temperature at whichLa phase signal first appears and the lowest temperature atwhich gel phase signal is no longer present as the ‘‘onset’’and the ‘‘end’’ of the transition, respectively. For the deuter-ated PSM sample, the transition takes place over 17�, fromonset at 39�C to the end at 56�C (both temperatures aremarked by boxes in Fig. 1). For the deuterated PCer sample,the transition width only spans 10� (47–57�C). The superpo-sition of gel and La phase signal in the transition regionmeans that the deuterated lipid is in both phases. The factthat, between 39 and 46�C, the 90:10 PSM-D31:PCerspectra contain some La phase while the 90:10PSM:PCer-D31 spectra do not, is clear evidence that purePSM-D31 La domains exist at these temperatures.

Average spectral width

The first moment, M1, is a measure of average spectralwidth. For a gel phase spectrum, M1 is approximately twicethat of an La spectrum. Fig. 2 shows the temperature depen-dence of M1 for all samples in a M1-T graph; gel-La phasetransitions manifest as segments having a large negative

slope. For each sample, the onset temperature, as deter-mined from spectral inspection, is within 6�C of the steepdrop and the end-transition temperature from inspectioncorresponds to the intersection between the steep and thehigh-temperature shallow M1-T segments. Increasing PCerconcentration has two notable effects on PSM:PCer MLVs:

1. The gel phase M1 of PCer (or PCer-D31) containingmembranes is noticeably increased compared to that ofpure PSM-D31.

2. Increased PCer content also increases the midpoint-temperature of the gel-La phase transition-temperaturefor both sphingolipids.

PCer, like other ceramides (37), enhances the gel phasestability of the membrane. For 80:20 PSM-D31:PCer and74:26 PSM-D31:PCer, the M1-versus-temperature curvesshowed plateaus in the vicinity of 53�C, likely due to fasterecho decay rates and concomitant changes to spectral shapeat these temperatures. Finally, we note that, at 40 mol %PCer, there is a prominent shoulder in M1 near 60

�C.

Liquid crystalline phase properties

In the La phase, each deuteron on the palmitoyl chaincontributes a Pake doublet to the spectrum. In general,deuterons of the same methylene group are equivalent andhave an average orientation perpendicular to the lipid longaxis. The deuterons on C2 and C3 of PSM-D31, however,have different orientations due to the structural kink ofthe amide-linked palmitoyl chain. Consequently, the

Biophysical Journal 103(12) 2465–2474

Page 4: Insights into Sphingolipid Miscibility: Separate Observation of Sphingomyelin and Ceramide N-Acyl Chain Melting

-10-8-6-4-20

1.2

1.0

0.8

0.6

0%

-10-8-6-4-20 1.4

1.2

1.0

0.8

0.6

10%

-6-4-20

x103

1.4

1.2

1.0

M1 (10

5s-1)

20%

2468 Leung et al.

temperature dependence of their quadrupolar splittings isreduced, allowing for their identification. We observe quad-rupolar splittings of 18 and 25 kHz for the C2 deuterons, and32 and 34 kHz for the C3 deuterons, in agreement withvalues given by Mehnert et al. (36) for oriented bilayersof selectively C2-labeled (19.4 and 28.9 kHz) and perdeuter-ated PSM (33 and 35 kHz for the C3 deuterons).

Having identified C2 and C3, order parameter profiles(Fig. 3) for PSM-D31:PCer and PSM:PCer-D31 membranesat 75�C can be calculated from resolved quadrupolar split-tings (Eq. 2), allowing detailed examination of the La phasesphingolipid membrane structure. The order parameters ofthe plateau region (C4–C9) are almost identical for thetwo sphingolipids at all concentrations, but the distal end(C10–C16) of PSM-D31 tends to be slightly more orderedthan PCer-D31. The order parameters of both PSM-D31and PCer-D31 increase with increasing PCer (or PCer-D31) concentration.

-10-8 0.8

0.6

-10-8-6-4-20 1.4

1.2

1.0

0.8

0.6

30%

0 1.4

Differential scanning calorimetry

Differential scanning calorimetry (DSC) thermogramsof Busto et al. (38) for nondeuterated PSM:PCer up to50 mol % PCer are analyzed in relation to 2H NMR results:Fig. 4 shows the negative integration curve of the DSC ther-mograms (dotted line) superimposed with 2H NMR M1-Tgraphs for PSM-D31:PCer (solid symbols) and PSM:PCer-D31 (open symbols), scaled to reflect relative changes. We

0.25

0.20

0.15

0.10

0.05

0.00

Ord

er P

aram

eter

161284

Carbon Number

161284

0.25

0.20

0.15

0.10

0.05

0.00

20%10%

30% 40%

FIGURE 3 Order parameter profiles for 90:10, 80:20, 70:30, and 60:40

PSM-D31:PCer (solid symbols) and PSM:PCer-D31 (open symbols) at

75�C. All membranes are in the La phase. SCD values within the plateau

regions are average values because the quadrupolar splittings from

deuterons on these carbons are unresolved.

-12

-8

-4

80706050403020Temperature (°C)

1.2

1.0

0.8

0.6

40%

FIGURE 4 The negative integral of the DSC thermograms of PSM:PCer

(dotted lines) are superimposed on 2H NMR M1 temperature-dependence

graphs for 100:0, 90:10, 80:20, 70:30, and 60:40 PSM-D31:PCer (solid

symbols) and PSM:PCer-D31 (open symbols). The two graphs are scaled

such that the maximum and minimum values of the integral and M1 are

aligned.

Biophysical Journal 103(12) 2465–2474

analyze the transitions at each concentration separately,starting with 2H NMR. For membranes containing<40 mol % PCer (or PCer-D31), the onset of the PCer-D31transition is higher than that for PSM-D31, signifying thatthe gel phase in the coexistence regions is enriched in PCer,relative to the overall sample composition. At 40%, PSM-D31 and PCer-D31 have similar ordering at all temperatures,indicating that the two lipids are well mixed and undergo thetransition simultaneously. From the jagged temperaturedependence ofM1 of 80:20 and 74:26PSM-D31:PCer (nomi-nally 30 mol % PCer), it can be deduced that PSM undergoescomplex phase behavior at these concentrations.

Page 5: Insights into Sphingolipid Miscibility: Separate Observation of Sphingomyelin and Ceramide N-Acyl Chain Melting

Insights into Sphingolipid Miscibility 2469

The negative integration curves of Fig. 4 are obtained byintegrating the DSC thermograms with respect to tempera-ture using the trapezoid method. The DSC traces (seeFig. S1 in the Supporting Material) show a relatively sharpendotherm at 40.9�C (0–30% PCer) associated with the gel-La transition of a pure PSM component; a broad transitionaround 70�C (40–50%), identified as the gel-La transitionof a homogeneous PSM:PCer mixture; and a plateaubetween these two transitions (2.5–10 mol % PCer). Theintegral of a DSC thermogram is proportional to enthalpy.The negative integral behaves like the 2H NMR M1 in thathigh values correspond to the gel phase, low values corre-spond to the La phase, and large negative slopes indicatephase transitions. Steeper drops indicate sharper transitions.For pure PSM, both the DSC integral and the 2H NMR M1

have the greatest slope at 40�C. For 10%, the integral tracesthe drop in the PSM M1. At 20 and 30%, however, the inte-gral more closely resembles the behavior of the PCer M1. At40%, the integral and the two M1 values drop sharply,evidence that a homogenous mixture of sphingolipids ispresent at all temperatures studied. Even though the DSCintegral is clearly more closely correlated with M1 inPSM-D31:Cer at 10%, and PSM:PCer-D31 at 20 and30%, DSC is a bulk technique and follows the cooperativethermal transitions of the whole sample. The two techniquesare in agreement in that all of the PSM and PCer meltingobserved by 2H NMR occurs at temperatures within thetransition region detected by DSC.

Phase diagram

Fig. 5 is a partial phase diagram that summarizes our DSCand 2H NMR findings. From DSC, the apex temperatureof the low-temperature endotherm (~40�C) for 0–30 mol %PCer, the low-temperature onset of the broad endotherm

80

70

60

50

40

30

Tem

pera

ture

(°C

)

50403020100

Cer Concentration (%)

PSM:PCer Lα

PSM GelPSM:PCer Gel

PSM:PCer LαPSM:PCer Gel

PSM LαPSM:PCer Gel

PSM:PCer Gel

FIGURE 5 Phase diagram constructed from spectral inspections of PSM-

D31:PCer (:), spectral inspections of PSM:PCer-D31 (6), spectral

subtractions of PSM:PCer-D31 (B), completion of the highest-temperature

DSC endotherm (-), and the apex temperature of the low-temperature

endotherm (~40�C) for 0–30% PCer and low-temperature onset of the

broad endotherm (~70�C) for 40 and 50% PCer on DSC thermograms (,).

(~70�C) for 40–50 mol % PCer, and the high-temperaturecompletion of the highest temperature endotherm for allPCer-containing samples are plotted. For 2H NMR, phaseboundaries were defined by spectral inspection. From 58to 61�C, phase boundaries were also inferred using spectralsubtraction of pairs of PSM:PCer-D31 spectra (see Methodsand Materials). The phase diagram consists of five regions,as follows.

Region 1. Low-temperature, low [PCer]: Below 40�C,membranes with %30 mol % PCer are composed oftwo coexisting gel phases—one of pure PSM andthe other a mixture of PSM and PCer.

Region 2. 40–50�C, low [PCer]: The pure PSMcomponent of membranes with %30 mol % PCermelts near 40�C, resulting in a pure PSM La

phase and a gel phase composed of PSM and PCerabove 40�C.

Region 3. Above 50�C, low [PCer]: Between ~50�C andthe liquidus, membranes containing 10–20 mol %PCer are composed of coexisting PSM:PCer La andgel phases.

Region 4. Low temperature, high [PCer]: A PSM:PCergel exists above 30 mol % PCer at low temperatures;there is no pure PSM component present.

Region 5. High-temperature: Above the liquidus, thetwo sphingolipids are in the La phase and well mixed.The liquidus temperature increases with increasingPCer content from 40�C for pure PSM to 73�C for50 mol % PCer.

These lipid behaviors are deduced in the following way.The sharp and invariant DSC endotherm near 40�C, inconjunction with the onset of PSM melting as seen inPSM-D31:PCer spectra and the complete lack of La signalin the PSM:PCer-D31 spectra, is compelling evidencefor a three-phase line that divides Regions 1 and 2. InRegion 3, an La component and a gel phase componentare evident in 2H NMR spectra for both PSM-D31:PCerand PSM:PCer-D31. In Region 4, there is no evidence ofa pure PSM gel phase: the 40�C endotherm is missingfrom DSC traces, and gel phase characteristics extend toconsiderably higher temperatures in 2H NMR spectra ofboth lipids. At 40 mol %, both DSC and 2H NMR indicatesthat the PSM and PCer are well mixed, forming a single gelphase that undergoes a gel-La transition between ~60 and72�C. At 50 mol %, the DSC peak sharpens further(Fig. S1). The liquidus that defines Region 5 is clearlydefined by both DSC and 2H NMR of membranes withlabeled PSM or PCer.

Theoretical ideal mixing of PSM:PCer in theLa phase

Let us assume that the free energy of interaction between thetwo sphingolipids is the same in a given phase and that PSM

Biophysical Journal 103(12) 2465–2474

Page 6: Insights into Sphingolipid Miscibility: Separate Observation of Sphingomyelin and Ceramide N-Acyl Chain Melting

2470 Leung et al.

and PCer mix ideally (that is, that the lipids will distributerandomly in each phase). Under such assumptions, the frac-tion of PCer at the liquidus and the solidus, xl and xg, respec-tively, at a given temperature, T, can be solved analyticallyusing regular solution theory (39):

xl ¼ eDG+PSM

=RT � 1

eDG+PSM

=RT � eDG+PCer

=RT; (5)

e�DG+PSM

=RT � 1

xg ¼

e�DG+PSM

=RT � e�DG+PCer

=RT: (6)

As usual, R is the molar gas constant and the free energyof melting is DG� ¼ DH� � TDS�. At the transition temper-ature, DG� ¼ 0, and DH� can be used to calculate DS�.Over a limited temperature range, DS� and DH� areapproximately constant. Fig. 6 shows the phase boundariescalculated from Eqs. 5 and 6 using known values ofDH

�PSM ¼ 4.35 � 104 J/mol, DH

�PCer ¼ 5.77 � 104 J/mol

(14) at the transition temperatures of TPSM ¼ 314.0 K (29)and TPCer ¼ 363.2 K (14). The experimental datafrom Fig. 5 are included again here for comparison: thecalculated liquidus matches the experimentally determinedliquidus, supporting the proposition that the two lipidsare well mixed in the PSM:PCer La phase up to atleast 50 mol % PCer. The calculated solidus, however,does not agree with the three-phase line at 40�C, providingfurther evidence that sphingolipid mixing is complex in thesolid phase.

80

70

60

50

40

30

Tem

pera

ture

(°C

)

50403020100

Cer Concentration (%)

FIGURE 6 Phase boundaries calculated from regular solution theory

based on the assumption of ideal mixing (solid lines) superimposed on

the experimental data of Fig. 5.

Biophysical Journal 103(12) 2465–2474

DISCUSSION

The conversion of SM to Cer is an important but not fullyunderstood process in the early stages of apoptosis. Togain insight into the nature of the physical changes thatthis transformation produces in the cell membrane, wefocused on a controlled model system composed of PSMand PCer. The major findings are

1. the presence of even a small amount of PCer affects theproperties of the PSM, and

2. the two lipids mix well in the liquid crystalline state, andfor high PCer contents, in the gel state.

Sphingomyelin-ceramide phase behavior

Fig. 7 portrays the essence of the phase diagram (Fig. 5).Two phase boundaries were determined with certainty:

1. an isothermal three-phase line at 41�C up to 30 mol %PCer, and

2. the liquidus that separates the PSM:PCer La phase regionfrom the region of phase coexistence below it for PCerconcentrations R10 mol %.

The three-phase line was indicated by a consistentlyobserved endothermic DSC transition and the appearanceof an La phase component in the 2H NMR data. The liquidus

FIGURE 7 The PSM:PCer partial phase diagram. (Solid lines) Liquidus

and the three-phase line at 41�C reflecting the good agreement of DSC and2H NMR data. Other phase boundaries are expected, and we offer possible

locations for these (dashed lines). An interpretation of the phases is also

shown and corresponds to labeled regions of the phase diagram in Fig. 5.

(Shaded and solid lipids) PSM and PCer, respectively. (Straight tails) Gel

phase lipids. (Curvy tails) La phase lipids.

Page 7: Insights into Sphingolipid Miscibility: Separate Observation of Sphingomyelin and Ceramide N-Acyl Chain Melting

Insights into Sphingolipid Miscibility 2471

was defined by the lowest temperatures at which no gelsignals were visible in 2H NMR spectra of both sphingoli-pids, the end-points obtained from 2H NMR spectralsubtraction, and the ends of the highest temperature DSCendotherms. Furthermore, all data fell close to the upperphase boundary predicted by theoretical calculations(Fig. 6). The liquidus at low PCer concentrations is deter-mined with less certainty, and indicated by a dotted line.

Between the three-phase line and the liquidus, thePSM:PCer system exhibits complex phase behavior (e.g.,Fig. 4) that is not fully understood. The phase boundariesin this region (Fig. 7, dotted lines) are inferred from theneed to separate different parts of the phase diagram butare less clearly indicated by the data. One of the three isan isothermal line shown at 53�C. Gibbs phase rule necessi-tates the existence of a three-phase line because it separatesregions of PSM:PCer LaþPSM:PCer gel coexistence andPSM La þPSM:PCer gel coexistence. Evidence supportingthis includes the appearance of La PCer, which happens at51 5 4�C for %30 mol % PCer, and the changes in slopein the M1-T graphs of PSM-D31:PCer near 53�C at 10,20, and 30% (Fig. 4). The conformational order of PSM’sN-linked chain in the La phase increases with increasingPCer concentration. This can be seen clearly in Fig. 2 forPSM at temperatures above the liquidus because M1 isdirectly proportional to the average value of the CD bondorder parameter (40). The incorporation of PCer into PSMis thus consistent with the observed reduction in the rateof change of M1 upon heating.

There are some discrepancies between the phasediagrams in Figs. 5 and 7. First, the deuteration of PSMmay lower the melting temperature of PSM slightly(41,42), which would explain why the liquidus determinedfrom spectral inspection of PSM-D31 spectra occursa degree or two lower than the predicted liquidus. Second,DSC shows a very small endothermic transition at 40�Cfor 70:30 PSM:PCer (Fig. S1); we interpret this as themelting of a tiny proportion of pure PSM. In contrast,NMR spectra for 74:26 PSM-D31:PCer show no liquid crys-talline signal below 47�C. Third, and finally, the phaseboundaries separating the single PSM:PCer gel phase regionfrom the neighboring mixed-phase regions are not preciselydetermined.

Sphingomyelin and ceramide in the gel phase

One striking feature of PSM:PCer membranes is that for allmixtures studied, PCer incorporates into the gel phasemembrane with no evidence of solid PCer formation. Thisis markedly different from, for example, POPC:PCer, wherePCer’s solubility limit is only z20 mol % (27,43). Incontrast, even at 40 mol %, the 2H NMR spectra for ourtwo sphingolipids are nearly identical for each temperaturein the gel phase, as reflected in the nearly superimposedM1-T plots (Fig. 4). DSC thermograms of the PSM:PCer

membrane, which display multicomponent endothermsfrom 2.5 to 30 mol % PCer (see Fig. S1 and Busto et al.(38)), become less complex at 40 and 50 mol % PCerwith no evidence from DSC of a pure PCer transitionclose to 90�C, the transition temperature for pure PCer inexcess buffer (14). The gel-to-La endotherm is significantlynarrower in the equimolar system than in 60:40 PSM:PCer;we speculate that the two sphingolipids form an azeotrope atsome composition close to equimolar.

Sphingomyelin and ceramide in the liquidcrystalline phase

PSM and PCer mix ideally in the liquid crystalline phase, asshown by coincident disappearance of gel phase sphingoli-pids at the liquidus and by the location of the liquidus on thephase diagram predicted by regular solution theory withideal mixing assumptions (Fig. 6). PCer concentrationdramatically affects PSM:PCer La phase formation: at10% PCer, the liquidus occurs at 56�C as opposed to the40�C of pure PSM. Thus, a small amount of PCer is suffi-cient to significantly enhance the ability of the membraneto exhibit some gel phase characteristics at elevated temper-atures. Despite mixing ideally in the La phase, PSM andPCer do not have identical chain order parameter profiles(Fig. 3): PSM is slightly more ordered, reflecting the influ-ence of the headgroup on the two sphingolipids’ intrinsicchain mobilities.

We know, however, that the N-linked palmitoyl chains ofPSM and PCer are kinked similarly because the C2deuterons of both sphingolipids are inequivalent. Note thatthe sphingolipid binary membrane is only slightly moreordered than POPC:PCer: 90:10 PSM:PCer has a maximumorder parameter of 0.225 at 75�C, whereas 90:10POPC:PCer has a maximum order parameter of 0.219 at57�C (27). In contrast, 1:1:1 PCer-D31:Cholesterol:Palmiticacid has a maximum order parameter of 0.43 at 50�C (43),reflecting the enhanced PCer chain order in the presence ofcholesterol.

Comparison of DSC and 2H NMR results

Fig. 4 is the first comparison of the integrated calorimetryresults with the temperature variation of the 2H NMR M1

done for a membrane composed of two lipids, each beingdeuterium-labeled (44–46). It is somewhat surprising inthat the DSC clearly tracks the changes in M1 exhibitedby the minor lipid component at 20 mol % PCer. Becauseof the dearth of similar published data, it is not knownwhether this is unusual. Linseisen et al. (46) used one deute-rium-labeled lipid (DPPC) in combination with a cationiclipid. DSC and NMR data for the major lipid component(DPPC) appeared to agree well, but when DPPC was only30% of the membrane, there were differences: the DSCwas multicomponent, whereas dM1/dT had only a single

Biophysical Journal 103(12) 2465–2474

Page 8: Insights into Sphingolipid Miscibility: Separate Observation of Sphingomyelin and Ceramide N-Acyl Chain Melting

2472 Leung et al.

peak. We know of no published examples where the minorcomponent in a binary mixture determines the bulk thermalbehavior of the membrane. However, we note that for20 mol % PCer, the integrated DSC does drop slightly at40�C, as the remaining pure PSM melts. The ensuingDSC integral’s tracking of the PCer-D31 NMR data implythat the PCer absorbs most of the thermal energy, in a coop-erative manner, between 45 and 65�C. 2H NMR shows thatwithin this temperature range, the PCer componentundergoes a gel to La transition.

Recent literature

Recent investigations into the phase behavior of modelmembranes containing Cer are numerous. We first describehow our work contributes to published data on the binaryPSM:PCer sphingolipid system. Busto et al. (38) studiedPSM:PCer monolayers and bilayers, and found that between2.5 and 30 mol % PCer, at least two Cer-containing phasesare present, along with a PSM phase. Fig. 7 shows that thecomposition of the two Cer-containing phases, a PSM:PCergel phase and a PSM:PCer liquid crystalline phase, varieswith temperature. Monolayers of PSM:PCer with 30–40 mol % PCer were identified as stable, composed of onlya liquid-condensed phase at 26�C (38). In MLVs, we showedthat the pure PSM phase is no longer present at R30 mol %PCer. Thus, in this region of the phase diagram there isa single PSM:PCer phase in monolayers and bilayers.

Fig. 8 of Busto et al. (38) shows room temperature giantunilamellar vesicles (GUVs) composed of PSM:PCer indifferent ratios labeled with the fluorescent probe DiIC18,which is purported to label PSM-enriched domains. Wepostulate, based on our results, that the PSM-enriched phasein Busto et al. (38) is composed of PSM and probe. Thedata are generally consistent with coexisting pure PSMand PSM:PCer gel phases. Qualitatively, the 2.5% PCerGUV has dark regions that cover less than one-quarter ofthe surface. This is consistent with the nearly verticalphase-boundaries at 0 and 32 mol % PCer in our phasediagram (Fig. 7): estimates using the lever rule suggestthat this membrane should display 8% dark domains. The10 mol % PCer GUV’s dark regions cover ~30% ofthe visible surface; our phase diagram predicts 31%. The20 mol % PCer GUV’s dark regions are ~50% of the visiblesurface, while our phase diagram’s prediction is 63%. At30 mol % PCer, dark regions dominate the surface of theGUV, consistent with our phase diagram’s predictionof <10% pure PSM. At 40 mol % PCer GUVs did notform. On the basis of the phase diagram, we expect tohave only the PSM:PCer gel phase present at this PCerconcentration. Our results are thus consistent with Bustoet al.’s conclusion that the PSM-enriched phase is neededfor GUV formation.

The GUV results agree remarkably well with the phasediagram especially given that inconsistencies might be ex-

Biophysical Journal 103(12) 2465–2474

pected due to differences among fluorescence microscopy,DSC, and 2H NMR. For example, especially for GUVshaving a small proportion of bright phase, the fluorescentprobe concentration in this bright phase becomes nonnegli-gible, e.g., ~5% probe (instead of the nominal 0.5%) in the30 mol % PCer GUV. It is likely that the higher probeconcentration will increase the amount of PSM-enrichedphase present. Despite the overall good agreement, careshould be taken as measurements of DiIC18-containingand DiIC18-depleted areas are only semiquantitative.Furthermore, the vesicles shown in Busto et al. (38) repre-sent three-dimensional reconstructions from half-GUVs.The contribution from the other half of the vesicle isunknown, and heterogeneous vesicle populations are thenorm in GUV electroformation.

More complex membrane systems

Knowledge of the phase diagram of a simple membrane suchas the PSM:PCer system studied here aids the understandingof more complex membrane systems. Such membranes canbe broadly categorized into those that have fixed composi-tion, and those that incorporate sphingomyelinase, andhence have variable sphingolipid concentrations.

For those membranes with fixed composition, Cer wasfound to have a high affinity for ordered domains, whethergel or liquid-ordered phase, in membranes composed ofequimolar SM:DOPC or DSPC:DOPC:cholesterol, respec-tively (47). Using fluorescence anisotropy of trans-parinaricacid, the formation of a PSM:PCer gel phase was observedupon addition of small amounts of PCer to 78:22PSM:POPC membranes (48). PCer also induces gel phaseformation in POPC:PCer in the absence of SM (27). Atomicforce microscopy yields additional information into Cerorganization within membranes: Cer-enriched gel phasesformed in equimolar mixtures of SM:DOPC:cholesterolupon the partial replacement of SM with Cer (31). Meghaand London (32) reported that Cer displaces sterol fromordered regions of heterogeneous membranes composed ofsaturated lipid (DPPC and Cer), low-melting lipid (spin-labeled PC or DOPC), and sterol (Chol and/or dehydroer-gosterol). In addition, these researchers reported that Cerstabilized ordered domains in the membrane, raising theirmelting temperature. These findings are all broadly consis-tent with formation of a stable PSM:PCer gel phase in ourbinary lipid mixture.

However, Cer-enriched gel phase formation associatedwith the displacement of cholesterol by Cer appears todepend on lipid composition. Adding cholesterol to 70:30PSM:PCer causes a homogeneous phase to form at roomtemperature (30), indicating that Cer, SM, and cholesterolcan all interact closely. This result highlights the importanceof understanding the interaction energies between pairs oflipids: a possible explanation for cholesterol’s unpredict-ability is that Cer can only displace cholesterol in an

Page 9: Insights into Sphingolipid Miscibility: Separate Observation of Sphingomyelin and Ceramide N-Acyl Chain Melting

Insights into Sphingolipid Miscibility 2473

environment that includes a cholesterol sink such as thatprovided by an unsaturated PC. Busto et al. (30) provided,as an alternative explanation, the stabilization of SM:Cer:Chol ternary phases, with no cholesterol displacement. Inprinciple, it would also be possible that Cer displacescholesterol to form cholesterol-dense regions, likely leadingto crystalline cholesterol monohydrate.

Several studies of the action of sphingomyelinase onSM in multicomponent synthetic membranes, mimickingthe early stages of apoptosis, have been published(24,31,49–51). Depending on the precise composition ofa membrane, enzyme action generally results in domainsthat consist of primarily SM and Cer. Necessarily, the localconcentration of Cer will vary considerably in thesemembranes. The knowledge of the structural predispositionsof PSM:PCer domains determined here gives insight into thephysical state of quasibinary sphingolipid domains in themore complex model membranes. In the absence of choles-terol, small angle x-ray scattering was used to monitor theaction of sphingomyelinase on POPC:egg SM comparedwith equilibrium mixtures of POPC:egg SM:PCer (51).PCer crystal formation occurred at high Cer concentrations(for example, 50:15:35 POPC:egg SM:PCer). Previously wefound a solubility limit of 20% PCer in POPC (27), but inthis article there is no evidence of PCer solid phase forma-tion. Adding POPC to the binary sphingolipid membranelikely reduces the solubility of PCer as observed (51).When cholesterol is included in the membrane, gel phasescan be generated by sphingomyelinase action, such as foundin DOPC:stearoylSM:Chol (31). Cer-enriched gel regionswere observed in regions of high enzyme activity. Basedon our phase diagram, we predict that these gel regionsare mainly composed of stearoylSM and stearoylCer.

CONCLUSION

The conversion of SM to Cer is key to several importantcellular processes: the early stages of apoptosis; internaliza-tion of membrane lesions needed for cell membrane repair(52); and the formation of the skin barrier, in particular atthe stratum granulosum/stratum corneum interface (53). Ifthe conversion is blocked, as in Niemann-Pick disease,abnormalities in membrane function ensue (54). Our studyshows conclusively that a Cer-containing gel phase formswith addition of even a small amount of Cer to SM; thus,presumably, the concentration of Cer found in early stageapoptotic plasma membranes is sufficient to cause physicalmodifications in SM-rich regions.

The lipid composition of the outer leaflet of mammalianplasma membranes is fairly well described by a simplemixture of SM, cholesterol, and a low-melting lipid suchas POPC. Sphingomyelinase action on this membranewould result in sphingolipid gel formation, provided thatthe system’s free energy is minimized when cholesterolassociates with the low-melting lipid. We have shown that

the gel phase formed by PSM and PCer is more stablethan the PSM gel phase, and that there is no PCer crystalformation up to at least 50 mol % PCer. It will be importantto further characterize the behavior of cholesterol interact-ing with SM and Cer to clarify whether, and how, ceramideand cholesterol compete.

SUPPORTING MATERIAL

DSC experiments and sample preparation, including one figure, are available

at http://www.biophysj.org/biophysj/supplemental/S0006-3495(12)01197-6.

We thank Joyce Leung for lipid phase illustrations.

J.T. and S.S.W.L. acknowledge the financial support of the Natural Sciences

and Engineering Research Council of Canada. F.M.G.’s work was sup-

ported by the Spanish Ministry of Economy (grant No. BFU 2007-62062)

and by the Basque Government (grant No. GIV 06/42). J.V.B. was sup-

ported by the Basque Government.

REFERENCES

1. Hannun, Y. A., C. R. Loomis, ., R. M. Bell. 1986. Sphingosineinhibition of protein kinase C activity and of phorbol dibutyratebinding in vitro and in human platelets. J. Biol. Chem. 261:12604–12609.

2. Kolesnick, R. N. 1987. 1,2-Diacylglycerols but not phorbol esters stim-ulate sphingomyelin hydrolysis in GH3 pituitary cells. J. Biol. Chem.262:16759–16762.

3. Hannun, Y. A., and L. M. Obeid. 2011. Many ceramides. J. Biol. Chem.286:27855–27862.

4. Lahiri, S., and A. H. Futerman. 2007. The metabolism and functionof sphingolipids and glycosphingolipids. Cell. Mol. Life Sci.64:2270–2284.

5. Goni, F. M., L. R. Montes, and A. Alonso. 2012. Phospholipases C andsphingomyelinases: lipids as substrates and modulators of enzymeactivity. Prog. Lipid Res. 51:238–266.

6. Stancevic, B., and R. Kolesnick. 2010. Ceramide-rich platforms intransmembrane signaling. FEBS Lett. 584:1728–1740.

7. Kolesnick, R. N., F. M. Goni, and A. Alonso. 2000. Compartmentaliza-tion of ceramide signaling: physical foundations and biological effects.J. Cell. Physiol. 184:285–300.

8. van Blitterswijk, W. J., A. H. van der Luit,., J. Borst. 2003. Ceramide:second messenger or modulator of membrane structure and dynamics?Biochem. J. 369:199–211.

9. Veiga, M. P., J. L. Arrondo, ., A. Alonso. 1999. Ceramides in phos-pholipid membranes: effects on bilayer stability and transition to non-lamellar phases. Biophys. J. 76:342–350.

10. Carrer, D. C., and B. Maggio. 1999. Phase behavior and molecularinteractions in mixtures of ceramide with dipalmitoylphosphatidylcho-line. J. Lipid Res. 40:1978–1989.

11. Cremesti, A. E., F. M. Goni, and R. Kolesnick. 2002. Role of sphingo-myelinase and ceramide in modulating rafts: do biophysical propertiesdetermine biologic outcome? FEBS Lett. 531:47–53.

12. Grassme, H., J. Riethmuller, and E. Gulbins. 2007. Biologicalaspects of ceramide-enriched membrane domains. Prog. Lipid Res.46:161–170.

13. Lofgren, H., and I. Pascher. 1977. Molecular arrangements of sphingo-lipids. The monolayer behavior of ceramides. Chem. Phys. Lipids.20:273–284.

14. Shah, J., J. M. Atienza,., G. G. Shipley. 1995. Structural and thermo-tropic properties of synthetic C16:0 (palmitoyl) ceramide: effect ofhydration. J. Lipid Res. 36:1936–1944.

Biophysical Journal 103(12) 2465–2474

Page 10: Insights into Sphingolipid Miscibility: Separate Observation of Sphingomyelin and Ceramide N-Acyl Chain Melting

2474 Leung et al.

15. Maggio, B. 1994. The surface behavior of glycosphingolipids inbiomembranes: a new frontier of molecular ecology. Prog. Biophys.Mol. Biol. 62:55–117.

16. Ramstedt, B., and J. P. Slotte. 2006. Sphingolipids and the formation ofsterol-enriched ordered membrane domains. Biochim. Biophys. Acta.1758:1945–1956.

17. Goni, F. M., A. Alonso,., J. L. Thewalt. 2008. Phase diagrams of lipidmixtures relevant to the study of membrane rafts. Biochim. Biophys.Acta. Mol. Cell. Biol. 1781:665–684.

18. Goni, F. M., and A. Alonso. 2009. Effects of ceramide and other simplesphingolipids on membrane lateral structure. Biochim. Biophys. ActaBiomembr. 1788:169–177.

19. Pinto, S. N., L. C. Silva, ., M. Prieto. 2011. Effect of ceramide struc-ture on membrane biophysical properties: the role of acyl chain lengthand unsaturation. Biochim. Biophys. Acta. 1808:2753–2760.

20. Metcalf, R., and S. A. Pandit. 2012. Mixing properties of sphingomye-lin ceramide bilayers: a simulation study. J. Phys. Chem. B. 116:4500–4509.

21. Holopainen, J. M., M. Subramanian, and P. K. Kinnunen. 1998.Sphingomyelinase induces lipid microdomain formation in a fluidphosphatidylcholine/sphingomyelin membrane. Biochemistry. 37:17562–17570.

22. Fanani, M. L., S. Hartel, ., B. Maggio. 2002. Bidirectional control ofsphingomyelinase activity and surface topography in lipid monolayers.Biophys. J. 83:3416–3424.

23. Hartel, S., M. L. Fanani, and B. Maggio. 2005. Shape transitions andlattice structuring of ceramide-enriched domains generated by sphingo-myelinase in lipid monolayers. Biophys. J. 88:287–304.

24. Fanani, M. L., S. Hartel, ., R. G. Oliveira. 2010. The action of sphin-gomyelinase in lipid monolayers as revealed by microscopic imageanalysis. Biochim. Biophys. Acta. 1798:1309–1323.

25. Sot, J., L. A. Bagatolli, ., A. Alonso. 2006. Detergent-resistant, ce-ramide-enriched domains in sphingomyelin/ceramide bilayers.Biophys. J. 90:903–914.

26. Huang, H. W., E. M. Goldberg, and R. Zidovetzki. 1996. Ceramideinduces structural defects into phosphatidylcholine bilayers andactivates phospholipase A2. Biochem. Biophys. Res. Commun. 220:834–838.

27. Hsueh, Y. W., R. Giles,., J. Thewalt. 2002. The effect of ceramide onphosphatidylcholine membranes: a deuterium NMR study. Biophys. J.82:3089–3095.

28. Bunge, A., P. Muller, ., D. Huster. 2008. Characterization of theternary mixture of sphingomyelin, POPC, and cholesterol: supportfor an inhomogeneous lipid distribution at high temperatures.Biophys. J. 94:2680–2690.

29. Maulik, P. R., and G. G. Shipley. 1996. N-palmitoyl sphingomyelinbilayers: structure and interactions with cholesterol and dipalmitoyl-phosphatidylcholine. Biochemistry. 35:8025–8034.

30. Busto, J. V., J. Sot, ., A. Alonso. 2010. Cholesterol displaces palmi-toylceramide from its tight packing with palmitoylsphingomyelin inthe absence of a liquid-disordered phase. Biophys. J. 99:1119–1128.

31. Chiantia, S., N. Kahya, ., P. Schwille. 2006. Effects of ceramide onliquid-ordered domains investigated by simultaneous AFM and FCS.Biophys. J. 90:4500–4508.

32. Megha, M., and E. London. 2004. Ceramide selectively displacescholesterol from ordered lipid domains (rafts): implications for lipidraft structure and function. J. Biol. Chem. 279:9997–10004.

33. Davis, J. H., K. R. Jeffrey, ., T. P. Higgs. 1976. Quadrupolar echodeuteron magnetic resonance spectroscopy in ordered hydrocarbonchains. Chem. Phys. Lett. 42:390–394.

34. Morrow, M. R., R. Srinivasan, and N. Grandal. 1991. The phasediagram of dimyristoyl phosphatidylcholine and chain-perdeuterateddistearoyl phosphatidylcholine—a deuterium NMR spectral differencestudy. Chem. Phys. Lipids. 58:63–72.

Biophysical Journal 103(12) 2465–2474

35. Vist, M. R., and J. H. Davis. 1990. Phase equilibria of cholesterol/dipal-mitoylphosphatidylcholine mixtures: 2H nuclear magnetic resonanceand differential scanning calorimetry. Biochemistry. 29:451–464.

36. Mehnert, T., K. Jacob, ., K. Beyer. 2006. Structure and lipid interac-tion of n-palmitoylsphingomyelin in bilayer membranes as revealed by2H-NMR spectroscopy. Biophys. J. 90:939–946.

37. Catapano, E. R., L. R. Arriaga,., I. Lopez-Montero. 2011. Solid char-acter of membrane ceramides: a surface rheology study of theirmixtures with sphingomyelin. Biophys. J. 101:2721–2730.

38. Busto, J. V., M. L. Fanani,., A. Alonso. 2009. Coexistence of immis-cible mixtures of palmitoylsphingomyelin and palmitoylceramide inmonolayers and bilayers. Biophys. J. 97:2717–2726.

39. Heimburg, T. 2007. Thermal Biophysics of Membranes. Wiley-VCHVerlag, Weinheim, Germany.

40. Davis, J. H. 1983. The description of membrane lipid conforma-tion, order and dynamics by 2H-NMR. Biochim. Biophys. Acta. 737:117–171.

41. Morse, R., L. D. Ma, ., F. Dunn. 1999. Ultrasound interaction withlarge unilamellar vesicles at the phospholipid phase transition: pertur-bation by phospholipid side chain substitution with deuterium. Chem.Phys. Lipids. 103:1–10.

42. Aussenac, F., M. Laguerre, ., E. Dufourc. 2003. Detailed structureand dynamics of bicelle phospholipids using selectively deuteratedand perdeuterated labels. 2H NMR and molecular mechanics study.Langmuir. 19:10468–10479.

43. Brief, E., S. Kwak, ., M. Lafleur. 2009. Phase behavior of anequimolar mixture of n-palmitoyl-d-erythro-sphingosine, cholesterol,and palmitic acid, a mixture with optimized hydrophobic matching.Langmuir. 25:7523–7532.

44. Morrow, M. R., J. C. Huschilt, and J. H. Davis. 1985. Simultaneousmodeling of phase and calorimetric behavior in an amphiphilicpeptide/phospholipid model membrane. Biochemistry. 24:5396–5406.

45. Morrow, M. R., and J. H. Davis. 1987. Calorimetric and nuclear-magnetic-resonance study of the phase-behavior of dilauroylphospha-tidylcholine water. Biochim. Biophys. Acta. 904:61–70.

46. Linseisen, F., S. Bayerl, and T. Bayerl. 1996. 2H-NMR and DSC studyof DPPC-DODAB mixtures. Chem. Phys. Lipids. 83:9–23.

47. Wang, T. Y., and J. R. Silvius. 2003. Sphingolipid partitioning intoordered domains in cholesterol-free and cholesterol-containing lipidbilayers. Biophys. J. 84:367–378.

48. Castro, B. M., R. F. M. de Almeida, ., M. Prieto. 2007. Formation ofceramide/sphingomyelin gel domains in the presence of an unsaturatedphospholipid: a quantitative multiprobe approach. Biophys. J. 93:1639–1650.

49. Ira, S., S. Zou, ., L. J. Johnston. 2009. Enzymatic generation ofceramide induces membrane restructuring: correlated AFM and fluo-rescence imaging of supported bilayers. J. Struct. Biol. 168:78–89.

50. Silva, L. C., A. H. Futerman, and M. Prieto. 2009. Lipid raft composi-tion modulates sphingomyelinase activity and ceramide-inducedmembrane physical alterations. Biophys. J. 96:3210–3222.

51. Boulgaropoulos, B., H. Amenitsch, ., G. Pabst. 2010. Implication ofsphingomyelin/ceramide molar ratio on the biological activity of sphin-gomyelinase. Biophys. J. 99:499–506.

52. Tam, C., V. Idone,., N. W. Andrews. 2010. Exocytosis of acid sphin-gomyelinase by wounded cells promotes endocytosis and plasmamembrane repair. J. Cell Biol. 189:1027–1038.

53. Kitson, N., J. Thewalt, ., M. Bloom. 1994. A model membraneapproach to the epidermal permeability barrier. Biochemistry.33:6707–6715.

54. Schuchman, E. H. 2010. Acid sphingomyelinase, cell membranes andhuman disease: lessons from Niemann-Pick disease. FEBS Lett.584:1895–1900.