Direct Observation of The Coexistence of Two Fluid Phases in Native Pulmonary Surfactant Membranes at Physiological Temperatures

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 9

Membrane Transport, Structure, Function,

and Biogenesis:
Cholesterol Rules: DIRECT
OBSERVATION OF THE
COEXISTENCE OF TWO FLUID
PHASES IN NATIVE PULMONARY
SURFACTANT MEMBRANES AT
PHYSIOLOGICAL TEMPERATURES

Jorge Bernardino de la Serna, Jesus Perez-Gil,


Adam C. Simonsen and Luis A. Bagatolli
J. Biol. Chem. 2004, 279:40715-40722.

Downloaded from http://www.jbc.org/ by guest on May 18, 2014


doi: 10.1074/jbc.M404648200 originally published online July 1, 2004

Access the most updated version of this article at doi: 10.1074/jbc.M404648200

Find articles, minireviews, Reflections and Classics on similar topics on the JBC Affinity Sites.

Alerts:
• When this article is cited
• When a correction for this article is posted

Click here to choose from all of JBC's e-mail alerts

This article cites 44 references, 6 of which can be accessed free at


http://www.jbc.org/content/279/39/40715.full.html#ref-list-1
THE JOURNAL OF BIOLOGICAL CHEMISTRY Vol. 279, No. 39, Issue of September 24, pp. 40715–40722, 2004
© 2004 by The American Society for Biochemistry and Molecular Biology, Inc. Printed in U.S.A.

Cholesterol Rules
DIRECT OBSERVATION OF THE COEXISTENCE OF TWO FLUID PHASES IN NATIVE PULMONARY
SURFACTANT MEMBRANES AT PHYSIOLOGICAL TEMPERATURES*
Received for publication, April 27, 2004, and in revised form, June 23, 2004
Published, JBC Papers in Press, July 1, 2004, DOI 10.1074/jbc.M404648200

Jorge Bernardino de la Serna‡, Jesus Perez-Gil‡, Adam C. Simonsen§, and Luis A. Bagatolli¶储
From the ‡Departmento de Bioquı́mica, Facultad de Biologı́a, Universidad Complutense, 28040 Madrid, Spain
and the MEMPHYS-Center for Biomembrane Physics, the §Department of Physics and the ¶Department of Biochemistry
and Molecular Biology, University of Southern Denmark, 5230 Odense, Denmark

Pulmonary surfactant, the lipid-protein material that lungs in air-breathing animals (1, 2). This surfactant material
stabilizes the respiratory surface of the lungs, contains is mainly composed of lipids, mainly phospholipids, and small
approximately equimolar amounts of saturated and un- amounts of specifically associated proteins. Among the phos-
saturated phospholipid species and significant propor- pholipids significant amounts of dipalmitoylphosphatidylcho-
tions of cholesterol. Such lipid composition suggests that line (DPPC)1 and phosphatidylglycerol are present, both of
the membranes taking part in the surfactant structures which are unusual species in most animal membranes. Mono-
could be organized heterogeneously in the form of in- unsaturated phosphatidylcholines (PC), phosphatidylinositol,
plane domains, originating from particular distributions and neutral lipids including cholesterol are also present in
of specific proteins and lipids. Here we report novel re- pulmonary surfactant in varying proportions. The surfactant

Downloaded from http://www.jbc.org/ by guest on May 18, 2014


sults concerning the lateral organization of bilayer mem- proteins (SP) SP-A, SP-B, SP-C, and SP-D constitute about 8%
branes made of native pulmonary surfactant where the
of lung surfactant by weight. SP-A and SP-D are large glyco-
coexistence of two distinct micrometer sized fluid phases
proteins (⬎500 kDa) belonging to the superfamily of collectins,
(fluid ordered and fluid disordered-like phases) is ob-
and they serve important functions in innate defense mecha-
served at physiological temperatures by using fluores-
cence microscopy and atomic force microscopy. Addi- nisms of the lung (3). The hydrophobic proteins SP-B and SP-C
tional experiments using fluorescent-labeled proteins modulate the surface-active properties of surfactant lipids and
SP-B and SP-C show that at physiological temperatures are strictly required to establish an operational respiratory
these hydrophobic proteins are located exclusively in the surface (4). Pulmonary surfactant lipid-protein complex is se-
fluid disordered-like phase. Most interestingly, the micro- creted into the thin aqueous alveolar lining of the lung as
scopic coexistence of fluid phases is maintained up to multilamellar assemblies, which spontaneously transform into
37.5 °C, where most fluid ordered phases melt. This obser- nanotubular membrane-based structures called tubular mye-
vation suggests that the particular composition of this lin. These structures are considered reservoirs of highly sur-
material is naturally designed to be at the “edge” of a face-active components in the pathway that forms surface-
lateral structure transition under physiological condi- active films at the lung air-water interface. Such films can
tions, likely providing particular structural and dynamic reduce the surface tension to nearly 0 mN/m and in doing so
properties for its mechanical function. The observed lat- prevent alveolar collapse at low lung volumes (1, 5). Dysfunc-
eral structure in native pulmonary surfactant membranes tions of the surfactant system are relevant in diseases includ-
is dramatically affected by the extraction of cholesterol, ing neonatal and acute respiratory distress syndrome, cystic
an effect not observed upon extraction of the surfactant fibrosis, and pneumonia (6). Several studies (7, 8) have shown
proteins. Furthermore, the spreading properties of the that interfacial films made with the hydrophobic fraction (lip-
native surfactant material at the air-liquid interface were ids ⫹ SP-B ⫹ SP-C) of pulmonary surfactant undergo phase
also greatly affected by cholesterol extraction, suggesting
separation under lateral compression. In addition, the partic-
a connection between the observed lateral structure and
ular phenomenon of gel/fluid phase separation induced by tem-
a physiologically relevant function of the material. We
perature was also reported in bilayers composed of part of the
suggest that the particular lipid composition of surfac-
tant could be finely tuned to provide, under physiological hydrophobic fraction of pulmonary surfactant at physiological
conditions, a structural scaffold for surfactant proteins to temperatures (9). Even though the particular composition of
act at appropriate local densities and lipid composition. the lung surfactant suggests that native surfactant membrane-
based structures could exhibit lateral segregation phenomena
at physiological temperatures, this aspect has been merely
Pulmonary surfactant is the main secretory product of the speculative up to now. For example, the presence of high
alveolar type II pneumocytes and is required to stabilize the amounts of DPPC (⬃40% weight of the total material) with a
melting phase transition above mammalian physiological tem-
peratures (41.5 °C) indicates the possibility of the coexistence
* This work was supported by Dirección General de Educación Supe-
rior e Investigaciones Cientı́ficas Grant BIO2003-09056 (to J. B. S. and of a solid/fluid-like phase (10). On the other hand, the presence
J. P.-G.), Comunidad Autónoma de Madrid Grant 08.2/0054, and by
Danish Natural Science Research Council Grant 21-03-0569 (to
1
L. A. B.). The costs of publication of this article were defrayed in part by The abbreviations used are: DPPC, 1,2-dipalmitoyl-sn-glycero-3-
the payment of page charges. This article must therefore be hereby phosphocholine; AFM, atomic force microscopy; Bodipy-PC, 2-(4,4-
marked “advertisement” in accordance with 18 U.S.C. Section 1734 difluoro-5,7-dimethyl-4-bora-3a,4a-diaza-s-indacene-3-pentanoyl)-1-
solely to indicate this fact. hexadecanoyl-sn-glycero-3-phosphocholine; DiIC18, 1,1⬘-dioctadecyl-
储 To whom correspondence should be addressed: MEMPHYS-Center for 3,3,3⬘,3⬘-tetramethylindocarbocyanine perchlorate; Me2SO, dimethyl
Biomembrane Physics, Dept. of Biochemistry and Molecular Biology, Uni- sulfoxide; DOPC, 1,2-dioleoyl-sn-glycero-3-phosphocholine; GUV, giant
versity of Southern Denmark, 5230 Odense M, Denmark. Tel.: 45-65-50- unilamellar vesicle; NPSM, native pulmonary surfactant material; SP,
34-76; Fax: 45-66-15-87-60; E-mail: bagatolli@memphys.sdu.dk. surfactant proteins; PC, phosphatidylcholine.

This paper is available on line at http://www.jbc.org 40715


40716 Phase Separation in Native Surfactant Membranes
of cholesterol along with other unsaturated phospholipid spe- viously (26). This procedure allows attachment of a single fluorophore to
cies in the native material suggests the possibility of a fluid the N-terminal amine group of the protein. In our studies Alexa 488 and
Texas Red were used for SP-C and SP-B, respectively. Labeled proteins
ordered/fluid disordered-like phase coexistence (distinguished
were incorporated into NPSM by addition of small aliquots of fluores-
in part by their cholesterol content) under physiologically rel- cent SP-B and SP-C (0.5% protein to lipid by weight) in Me2SO.
evant conditions (10). Even though lung surfactant has choles- Giant Vesicle Preparation—GUVs composed of NPSM organic ex-
terol concentrations of up to 20 –22 mol % (11, 12), there is no tract, its lipid fraction, or the ternary mixture DOPC/DPPC/cholesterol
clear understanding of how this molecule impacts the lateral were prepared as described previously (16, 27) by using the electrofor-
structure of the native material. The presence of cholesterol mation method originally developed by Angelova and Dimitrov (28) and
Angelova et al. (29). A previously described (16, 27) special tempera-
could thus have important consequences for the lateral organi-
ture-controlled chamber was used for this purpose. Briefly the process
zation of the native pulmonary surfactant material, illustrat- can be described in the following three steps. 1) ⬃3 ␮l of the stock
ing another example of the fundamental role of cholesterol in solution of the lipid or lipid/protein organic solution were spread on the
modulating the lateral structure of membranes, as the raft surface of each platinum wire. The chamber was located under a stream
hypothesis proposes (13, 14). Furthermore, there is presently of N2 during this procedure and then placed under vacuum overnight to
no clear link between membrane lateral heterogeneity and the remove the organic solvent. 2) Aqueous solution was added to the
chamber (200 mosM sucrose solution prepared with Millipore-filtered
physiological function of native pulmonary surfactant material
water, 17.5 megohms/cm). The sucrose solution was previously heated
(10). to the desired temperature (above the lipid mixture phase transition,
In recent years, a new experimental strategy, based on the we used 45 °C), and then sufficient volume was added to cover the
direct visualization of free standing lipid bilayers using giant platinum wires (⬃300 ␮l). 3) The platinum wires were connected im-
unilamellar vesicle technology in conjunction with confocal and mediately to a function generator (Digimess® FG 100, Germany), and a
two photon excitation fluorescence microscopy techniques, has low frequency AC field (sinusoidal wave function with a frequency of 10
Hz and an amplitude of 3 V) was applied for 120 min. After vesicle
opened the possibility of correlating morphological and dynam-
formation, the AC field was turned off, and the vesicles were collected
ical information on lipid membranes at molecular and supra- with a pipette and transferred to a plastic tube.
molecular levels. Important novel information such as the mor-

Downloaded from http://www.jbc.org/ by guest on May 18, 2014


NPSM GUVs were also prepared using the electroformation method
phology of different coexisting lipid phases (such as gel/fluid but with some modifications in step 1 of the protocol described above. In
and fluid ordered/fluid disordered), mechanical properties of this case NPSM GUVs were formed from NPSM suspended in buffer
membranes displaying phase coexistence, local hydration, and solution (no organic solvents in this case). Briefly, ⬃3 ␮l of the stock
suspension in buffer was spread on the surface of each platinum wire as
molecular diffusion in lipid domains can be extracted directly
small drops. The chamber was then placed under a stream of N2 and
from the fluorescence images (9, 15–23). The present study subsequently under low vacuum for 30 min to allow the native material
uses the GUV technology in addition to other experimental to adsorb onto the platinum wire. An important point in this step is to
techniques to provide evidence that a particular lateral struc- avoid dehydration of the sample to maintain the integrity of the native
ture occurs in native membranes of pulmonary surfactant at membranes. Once the material was adsorbed to the Pt wire, we proceed
physiological temperatures. Additionally we demonstrate that with steps 2 and 3 as described above. As a control experiment, GUVs
composed of DOPC/DPPC/cholesterol were also prepared from multila-
cholesterol plays a critical role in promoting such membrane
mellar suspensions in aqueous solution as described above for NPSM
organization. On the other hand, it is important to note that and then compared with those obtained from the lipid organic extract.
preparation of GUVs was previously limited to membranes No differences in the morphology of the final resulting GUVs were
composed of artificial and natural lipid extracts (from lipid observed between the two described protocols.
solutions in organic solvents). In this study we also present a Observation of Giant Vesicles—Aliquots of giant vesicles suspended
method to prepare GUVs composed of native membranes with- in sucrose were added to an equi-osmolar concentration of glucose
solution. Because of the density difference between the two solutions,
out using organic solvents or detergents. To our knowledge,
the vesicles precipitate at the bottom of the chamber which facilitates
GUV experiments performed with native membrane prepara- observation of the GUVs in the inverted confocal microscope. GUV
tions to further evaluate the concurrent effect of lipids and preparations were observed in 8-well plastic chambers (Lab-Tek Brand
proteins in lateral segregation phenomena have not yet been Products, Naperville IL). The chamber was located in an inverted
reported. confocal microscope (Zeiss LSM 510 META) for observation. The exci-
tation wavelengths were 488 (for Alexa 488 and Bodipy-PC) and 543 nm
EXPERIMENTAL PROCEDURES (for DiIC18 and Texas Red). The temperature was controlled from a
Materials—1,1⬘-Dioctadecyl-3,3,3⬘,3⬘-tetramethylindocarbocyanine water bath connected to a homemade device into which the 8-well
perchlorate (DiIC18), 2-(4,4-difluoro-5,7-dimethyl-4-bora-3a,4a-diaza-s- plastic chamber was inserted. The temperature was measured inside
indacene-3-pentanoyl)-1-hexadecanoyl-sn-glycero-3-phosphocholine the sample chamber using a digital thermocouple (model 400B, Omega
(Bodipy-PC), and isothiocyanate derivatives of Alexa Fluor® 488 and Inc., Stamford, CT) with a precision of 0.1 °C.
Texas Red® were from Molecular Probes Inc. (Eugene, OR). Methyl-␤- Differential Scanning Calorimetry—Experiments were performed in
cyclodextrin was from Aldrich. DPPC, 1,2-dioleoyl-sn-glycero-3- a Science Corp. N-DSCII calorimeter using 15 mM total lipid concen-
phosphocholine (DOPC), and cholesterol were purchased from Avanti tration of native pulmonary surfactant suspension. The scan rate was
Polar Lipids (Alabaster, AL) and were used without further purifica- 0.5 K/min versus buffer. Five temperature scans were collected from
tion. Native pulmonary surfactant material was obtained from pig each sample between 25 to 70 °C. Neither change in the total concen-
lungs as described previously (24). The cholesterol quantitation kits tration nor in the scan rate gave significant changes in the obtained
INFINITYTM and IVD were purchased from Sigma and Spinreact results.
(Girona, Spain), respectively. Cholesterol Extraction Experiments—Methyl-␤-cyclodextrin/glucose
Preparation of Surfactant and Proteins—Stock suspensions of native solutions were prepared in Millipore water 17.5 megohms/cm. The
pulmonary surfactant material (NPSM) were in 5 mM Tris buffer, pH 7, osmolarity of these solutions was measured in an osmometer (The
containing 150 mM NaCl. The total concentration of phospholipid was Advance Osmometer, model 3D3, Advance Instruments Inc., Norwood,
0.5 mg/ml, estimated by phosphorus quantitation upon phospholipid MA) and adjusted to match the osmolarity of the vesicles containing
mineralization (25). For fluorescence experiments the fluorescent sucrose solution. Vesicles were diluted in the methyl-␤-cyclodextrin/
probes were incorporated into NPSM stock suspensions before prepar- glucose-containing solution and observed afterward. Alternatively, ali-
ing the giant vesicles. In the case of the lipophilic probes (DiIC18 and quots of methyl-␤-cyclodextrin solutions were added to the vesicles
Bodipy-PC) a small aliquot (0.5 ␮l) of an Me2SO solution containing the already dispersed in glucose solution. No differences in the effect of
fluorescent probes (0.125 mM) was added to the NPSM stock suspension methyl-␤-cyclodextrin on the vesicles were observed between the two
(final volume 500 ␮l) and incubated for 20 min at room temperature. In procedures. Cholesterol was also removed from the lipid fraction of
all cases the percentage of fluorescent probes in the sample was less NPSM by hydrophobic chromatography in Sephadex LH-20 (30).
than 0.1 mol %. Phospholipid and Cholesterol Determination—Total phospholipid
Fluorescently labeled surfactant proteins were prepared by labeling concentration in the different samples was determined by phosphorus
purified SP-B and SP-C solutions in organic solvent, as described pre- analysis as we described above (25). The amount of cholesterol in NPSM
Phase Separation in Native Surfactant Membranes 40717
(15, 19), we conclude that the lateral organization of the NPSM
vesicles corresponds to a fluid ordered/fluid disordered-like
phase coexistence. The coexistence of two fluid phases persists
from 22 to 37.5 °C. In general, when fluid domains are embed-
ded in a fluid environment, circular domains are formed be-
cause both phases are isotropic, and the line energy (tension),
which is associated with the rim of two demixing phases, is
minimized by optimizing the area-to-perimeter ratio. As ob-
served in Fig. 1C, the domains span the bilayer in agreement
with previous observations done in artificial and natural lipid
FIG. 1. Giant vesicles from native pulmonary surfactant mem- mixtures (15–18). NPSM frequently gave rise to multilamellar
branes. Giant unilamellar (A–C) and multilamellar (D) vesicles formed giant vesicles (see Fig. 1D), whose morphology resembles the
from native pulmonary surfactant material. A, transmission image; structure of lamellar bodies that are the multilamellar assem-
B–D, fluorescent images. The confocal planes in the fluorescent images blies storing and secreting pulmonary surfactant in the type II
were obtained at the polar (B) and center (C and D) regions of the giant
vesicles. The temperature was 37 °C. The giant vesicles are labeled with pneumocytes of the lung. Coexistence of two fluid phases was
the lipophilic fluorescence probes DiIC18 (red) and Bodipy-PC (green). In observed in every bilayer of these giant multilamellar assem-
the NPSM the rounded green areas in the images correspond to fluid blies. Fig. 2 shows an AFM image of an NPSM-supported
disordered regions, and the red background corresponds to fluid ordered membrane, without exogenous fluorescent probes. Such mem-
phase. The scale bars correspond to 10 ␮m. Note that the different
domains span the lipid bilayer (C and D). branes are part of the surface-active film of pulmonary surfac-
tant that spontaneously forms upon adsorption at the interface
and in their different fractions was quantitatively estimated using the (32). It is important to emphasize that the samples visualized
following: (i) a colorimetric assay based on the activity of cholesterol by AFM are not simple monolayers, as those studied previously

Downloaded from http://www.jbc.org/ by guest on May 18, 2014


oxidase (IVD kit from Spinreact, Girona), and (ii) the determination of in the literature, but bilayers as we perform a double deposition
peroxide products resulting from cholesterol oxidation (INFINITYTM of surfactant films onto mica (see “Experimental Procedures”).
kit from Sigma). Cholesterol content in the samples is given as choles- AFM scanning of supported membranes demonstrates the
terol-to-phospholipid molar ratio. Four different samples were used for
total cholesterol quantitation from each of at least three different
presence of circular-shaped domains with sizes comparable
batches of each of the fractions studied. with those seen in GUVs, confirming that two fluid phases
Protein Extraction from NPSM—The hydrophobic fraction of NPSM, coexist in the native material. This observation represents an
containing all the lipid species plus the hydrophobic proteins SP-B and important control experiment because it indicates that the
SP-C but lacking hydrophilic proteins SP-A and SP-D, was obtained by membrane lateral structure in the GUVs is really present in
chloroform/methanol extraction. Separation of lipid from protein com-
the original material and is not because of the incorporation of
ponents in the organic mixture was achieved by chromatography of the
surfactant organic extract through a Sephadex LH-20 column, as de- fluorescent probes or the way the samples are prepared. AFM
scribed previously (31). also reveals the coexistence of nano-scale structures inside the
Atomic Force Microscopy of NPSM Bilayers—NPSM aqueous suspen- circular-shaped domains. Topological analysis shows that the
sions were spread on top of an ultrapure water subphase in a Langmuir- distribution of heights observed in the round domains (i.e.
Blodgett trough (Kibron, ␮-trough) until a surface pressure of 1 mN/m higher versus lower heights) is more heterogeneous than the
was obtained. The film was allowed to equilibrate for 10 min and was
subsequently compressed to 42 mN/m. While maintaining this pressure,
height distribution observed in the more homogeneous inter-
a film was transferred to a freshly cleaved mica substrate previously vening background (Fig. 2). In particular, the round domains
immersed into the subphase by lifting the support at a constant speed contain regions of lower heights compared to the height ob-
of 1.5 mm/min. Following the initial deposition, a second layer was served in the continuous region. This last finding suggests that
transferred onto the first by re-immersion of the coated mica into the 1) the round domains may correspond to a liquid disordered-
subphase at 1.5 mm/s (also at a constant surface pressure of 42 mN/m).
like phase state and that 2) the region surrounding the round
Consequently, the supported membranes used in the atomic force mi-
croscopy (AFM) experiments are bilayer structures formed by double domains may display a liquid ordered-like phase state. The
transfer of the surface film (including any additional associated struc- higher heights in the round domains may be due to the pres-
ture to the monolayer), which is obtained by spreading aqueous sus- ence of surfactant proteins. The complex topography of the
pensions of native surfactant at the air-water interface. Topographical liquid-disordered domains is likely connected with particular
AFM images were obtained under aqueous conditions in a PicoSPM organization of the many different molecular species (lipids
(Molecular Imaging) microscope operated in magnetically activated tap-
ping mode using MAClevers.
and proteins) in the plane of the membrane. Given the compo-
Spreading Kinetics—Surface activity of surfactant preparations was sitional complexity of the material, further experiments to clar-
assayed by running ␲-t interfacial adsorption isotherms. In a typical ify this issue are warranted but are outside of the scope of the
experiment, 3 ␮l of surfactant suspension (10 mg/ml phospholipid con- present study.
centration) were spread by direct deposition on top of a subphase 5 mM The particular domain pattern observed on the GUV surface
Tris, pH 7, containing 150 mM NaCl in the Teflon trough (20 cm2) of a
is likely to be relevant for modulating the lateral distribution of
thermostated surface balance built by Nima (Coventry, UK), and at 5
cm of a filter paper Wilhelmy plate, before monitoring surface pressure the surfactant hydrophobic proteins SP-B and SP-C in both the
against time. Spreading experiments were done either at 25 or 37 °C. native membranes and those formed from surfactant organic
Spreading experiments in the presence of methyl-␤-cyclodextrin were extract, as shown in Fig. 3. When fluorescently labeled SP-B
carried out by spreading equivalent amounts of native surfactant on top and SP-C were incorporated into NPSM GUVs or into GUVs
of 5 mM Tris, pH 7, subphases containing 150 mM NaCl, and different formed from the hydrophobic fraction of NPSM, the fluores-
concentrations of methyl-␤-cyclodextrin. Data shown are representative
of three different experiments for two different batches of each assayed
cence of these proteins colocalized with the Bodipy-PC-enriched
surfactant preparation. areas (from 22 to 37.5 °C). This last fact indicates that the
hydrophobic proteins segregate into fluid disordered regions, in
RESULTS agreement with the observations obtained in the AFM experi-
Fig. 1 shows images of single GUVs composed of NPSM ments (Fig. 2). The partitioning of SP-B and SP-C in NPSM and
doped with the fluorescent probes Bodipy-PC and DiIC18. in bilayers made of surfactant organic extract resemble the
Based on the particular round shape of the domains (18, 20, 21) distribution of the proteins observed in monolayer experi-
and the partition properties of the different fluorescent probes ments, where both SP-B and SP-C preferentially situate in
40718 Phase Separation in Native Surfactant Membranes

FIG. 2. AFM scanning image from a


pulmonary surfactant membrane. A,
AFM scanning image of membranes
formed by spreading NPSM on top of an
air-liquid interface, at a lateral pressure
of 42 mN/m. Film transfer was performed
using the Langmuir-Blodgett method
onto a mica support. The supported mem-
brane was scanned in a liquid cell using
tapping mode. B, magnification of the re-
gion marked in A. C, topographic profile
along the white line included in B. The
temperature was 25 °C.

Downloaded from http://www.jbc.org/ by guest on May 18, 2014


liquid ordered phase has melted, yielding less than 10% of the
surface covered by a network of filamentous shapes, probably
ordered-like phase. Fig. 4B shows a representative differential
scanning calorimetry experiment obtained from NPSM solu-
tions. The thermogram is characterized by a broad thermal
transition showing the end of the melting process above
37.5 °C, in agreement with the data obtained in GUVs and that
previously reported for alveolar surfactant (crude lung wash)
dispersion from mongrel dogs (35). Additionally, a very similar
effect is observed by using fluorescent-labeled proteins SP-B
and SP-C instead of the lipophilic fluorescent probes. From our
experimental data we can observe that the proteins remain in
the fluid disordered areas at temperatures above the lateral
structure change as observed at 37.5 °C (Fig. 4C).
The similarities between the fluorescent images obtained in
the NPSM and those reported previously (18, 20 –22) in choles-
FIG. 3. Lateral distribution of surfactant proteins SP-B and terol-containing ternary lipid mixtures are remarkable and
SP-C in pulmonary surfactant membranes. Fluorescence images of
GUVs composed of NPSM (A) or NPSM organic extract (B) upon incor- suggest a potentially important role for cholesterol in the lat-
poration of fluorescently labeled proteins SP-B and SP-C (0.5% protein/ eral structural organization of NPSM. However, because of the
lipid by weight). The two color images (bottom images) are split into compositional complexity of NPSM, the potential influence of
those corresponding to Alexa 488-labeled SP-C (green) and Texas Red- other components, in particular surfactant proteins, on the
labeled SP-B (red). Scale bars are 10 ␮m. Images were taken at 37 °C.
lateral arrangement of this membranous system had to be
liquid-expanded areas (33). This distribution is particularly evaluated. To address this point we used different experimen-
remarkable in the case of SP-C, because this protein should, in tal strategies involving the extraction of particular components
principle, have a different local environment in bilayers with of NPSM. Cholesterol extraction with methyl-␤-cyclodextrin
respect to monolayers simply because of the different mem- and extraction of surfactant proteins (SP-A, SP-B and SP-C)
brane thickness (SP-C in bilayers accommodate in a transmem- were performed on the NPSM, and the results are summarized
brane orientation (1)). Our results seem also to discard the in Fig. 5A. Neither the absence of the water-soluble fraction
potential association of SP-C with cholesterol-enriched liquid (including the collectin SP-A) nor the extraction of the hydro-
ordered regions of the membranes, in contrast with that de- phobic proteins SP-B and SP-C from the NPSM changed the
scribed for other palmitoylated transmembrane proteins (34). circular shape of the lipid domains between 22 and 37.5 °C.
Additionally, we observed that removal of SP-A from NPSM This result indicates that the presence of the surfactant pro-
does not affect the partition properties of either SP-B or SP-C teins is not critical to maintain the phase coexistence pattern
proteins (Fig. 3B). observed in NPSM. The cholesterol-to-phospholipid molar ratio
Fig. 4 illustrates the effect of temperature on the lateral in all these materials was similar (between 18 and 21 mol %),
structure of NPSM. The liquid disordered domains showed as summarized in Table I. Fig. 5A also shows that after iso-
dynamically fluctuating borderlines when the temperature was thermal cholesterol extraction with methyl-␤-cyclodextrin, the
raised close to 37.5 °C followed by a dramatic lateral structure phase coexistence pattern characterized by round domains in
change above 37.5 °C. Fig. 4A shows that at 38 °C, most of the NPSM changes to elongated irregularly shaped domains, sim-
Phase Separation in Native Surfactant Membranes 40719

FIG. 4. Effect of temperature on the


structure of native pulmonary sur-
factant membranes. Representative
confocal fluorescence images of giant
unilamellar vesicles composed of native
pulmonary surfactant material at the in-
dicated temperatures obtained with the
lipophilic fluorescent probes DiIC18 and
Bodipy-PC (A) and fluorescently labeled
SP-B and SP-C (C). The fluorescent image
in C is split into two colors corresponding
to Alexa 488-labeled SP-C (green) and
Texas Red-labeled SP-B (red). Scale bars
are 20 ␮m. Differential scanning calorim-
etry thermogram of NPSM (B).

Downloaded from http://www.jbc.org/ by guest on May 18, 2014


ilar to those reported previously in artificial ternary mixtures tration of methyl-␤-cyclodextrin (compare Fig. 5, A and C).
at very low cholesterol levels (20 –22) and in DPPC-containing Increasing the cholesterol concentration to 15 mol % induces
phospholipid binary mixtures (15, 16). The lateral structure the appearance of numerous circular-shaped domains. Further
observed in these systems corresponds to the coexistence of gel increase in cholesterol (up to 30 mol % relative to phospholip-
and fluid phases. To explain further the observed cholesterol- ids) yields an increase in the area occupied by the fluid-ordered
dependent phase changes in NSPM at constant temperature, phase (Fig. 5C, red area in the GUVs containing 15–30 mol %
cholesterol-extraction experiments using methyl-␤-cyclodex- cholesterol). Note that size and morphology of the fluid disor-
trin were performed in the well characterized DOPC/DPPC/ dered domains become comparable with those observed in
cholesterol mixture (20 –22). This particular lipid mixture was NPSM (Fig. 5A) upon addition of 8 –10 mol % of cholesterol,
chosen as a model system because the presence of DPPC and which raises the cholesterol content to values in the same
cholesterol is similar to that of NPSM. Fig. 5B shows two range as those in native pulmonary surfactant. From this ex-
different phase transitions as a result of increasing methyl-␤- periment, we estimate that a cholesterol-to-phospholipid ratio
cyclodextrin concentration at constant temperature, i.e. fluid of about 13–15 mol % triggers a phase change from gel/fluid to
ordered 3 fluid ordered/fluid disordered and fluid ordered/fluid fluid ordered/fluid disordered phase coexistence. These exper-
disordered 3 gel/fluid. The last phase transition occurs in a iments clearly demonstrate that the cholesterol concentration
remarkably similar fashion to that observed in NPSM (com- is a critical parameter in modulating the lateral structure of
pare Fig. 5, A and B). The domain compositional information pulmonary surfactant membranes. Moreover, the cholesterol-
already reported for DOPC/DPPC/cholesterol in the fluid or- dependent phase coexistence pattern reported here may ex-
dered/fluid disordered phase coexistence regime (20 –22) sug- plain why the phase coexistence pattern reported previously in
gests that the DPPC-containing fluid ordered phase is enriched GUVs composed of the bovine lipid extract surfactant BLES®,
in cholesterol compared with the DOPC-enriched fluid disor- a clinical preparation, corresponds to gel/fluid phase coexist-
dered phase where the cholesterol content is low. Based on ence (9). BLES® is almost devoid of cholesterol, and in this
these observations we propose that the fluid ordered phase in respect it is similar to NPSM at relatively high methyl-␤-
NPSM may be enriched in DPPC and cholesterol compared cyclodextrin concentration.
with the fluid disordered phase. Additionally, from our exper- One key question is whether the lateral structure of the
iments we notice that although the shape of the lipid domains surfactant material may affect its functional properties. Rapid
is quite similar for NPSM and the cholesterol-containing ter- spreading and adsorption of lung surfactant material to form
nary mixtures in the same phase coexistence regime (compare surface-active films at the air-water interface is an important
Fig. 5A, top left image, with B, center image, and Fig. 5A, property that can be used to evaluate the proper function of a
bottom left image, with B, bottom image), there are some dif- given surfactant. Fig. 6 presents spreading-at-interface ␲-t iso-
ferences in the partition properties of the fluorescent probes. therms of different surfactant suspensions directly deposited
This observation suggests that the molecular composition of on top of open buffered subphases. Both native surfactant as
these membrane domains (i.e. fluid ordered, gel, fluid disor- purified from porcine lungs and its reconstituted organic frac-
dered) could play an important role in modulating specific tion (lipids ⫹ SP-B ⫹ SP-C) have optimal adsorption activities,
molecular interactions in the plane of the membrane in agree- very rapidly producing equilibrium surface pressures around
ment with previous results observed in lipid mixtures (18). 45 mN/m (Fig. 6, A and B). Suspensions lacking proteins show
We found that the changes observed as the cholesterol con- much lower adsorption rates and do not reach surface pres-
centration is decreased (Fig. 5A) can be reversed by re-intro- sures higher than 30 mN/m after 50 min, indicating the essen-
ducing cholesterol as shown in Fig. 5C. GUVs formed from tial role of hydrophobic proteins to catalyze phospholipid inter-
surfactant lipid extract partially depleted of cholesterol (cho- facial transference. Most interestingly, spreading isotherms of
lesterol to phospholipid ratio 10 mol %, instead of 20 mol % as suspensions prepared from the surfactant lipid fraction par-
in NPSM) show a similar pattern of gel/fluid phase coexistence tially depleted of cholesterol are further impaired, suggesting
as that observed in NPSM GUVs treated with a high concen- that cholesterol-containing surfactant structures have some-
40720 Phase Separation in Native Surfactant Membranes
cyclodextrin in the subphase (i.e. the higher amount of choles-
terol extracted from the spreading surfactant), the lower the
maximum pressure reached at the late spreading stages. Op-
timal insertion of phospholipids against high pressures seems
therefore to require the action of surfactant-specialized struc-
tures such as the one described here when both cholesterol and
hydrophobic proteins are present.

DISCUSSION
The present study offers a novel view on the importance of
cholesterol for the relationships in the structure-function of
pulmonary surfactant, and suggests a rationale for the pres-
ence of a significant amount of this lipid species in the compo-
sition of all the known surfactants. In this paper we demon-
strate that the presence of cholesterol promotes a defined
lateral structure in surfactant bilayers at physiological temper-
atures. The particular lateral organization of lipids and pro-
teins in surfactant membranes would be essential to support
not only a rapid interfacial adsorption to equilibrium surface
pressures (as shown in the Fig. 6), but for the formation of well
defined surface-associated structures (such as that shown in
the Fig. 2). This last scenario would be competent to support
proper dynamic surface behavior when the surfactant material

Downloaded from http://www.jbc.org/ by guest on May 18, 2014


is subjected to rapid dynamic cycling.
Several studies have analyzed previously the effects of cho-
lesterol on the surface activity of surfactant, mostly using
model lipid mixtures. However, the conclusions about a poten-
FIG. 5. Fluorescent images of GUVs made from different sur- tial role of cholesterol in surfactant function have been contra-
factant fractions and lipid mixtures. A, GUV composed of NPSM
(top left) and GUV composed of the organic extract of NPSM that dictory. Some studies have concluded that cholesterol has a
contains all the hydrophobic components but no hydrophilic proteins general beneficial effect on surfactant interfacial adsorption,
(top right); GUV composed of the lipid fraction of the NPSM organic mostly due to a “fluidizing” effect on DPPC-enriched surfactant
extract, from which hydrophobic proteins SP-B and SP-C have also been lipid systems (36). On the other hand, the surface behavior
removed (bottom right), and GUV of NPSM treated with 40 mg/ml
methyl-␤-cyclodextrin (bottom left). Images were taken at 25 °C. The
studies of cholesterol-containing monolayers using model lipid
cholesterol-to-phospholipid molar ratio for the samples that display mixtures has come to the conclusion that cholesterol impairs
coexistence of two fluid phases (round-shaped domains) are 21% (top left the ability of any lipid mixture to reach and sustain very high
image), 19% (top right image), and 18% (bottom right). B, fluorescence surface pressures during compression (37–39). As a conse-
images of GUVs composed of DOPC/DPPC/cholesterol (1:1:40 mol %)
quence of these studies, many of the clinical surfactants used
before (top) and after exposure to 10 (center) and 40 mg/ml (bottom) of
methyl-␤-cyclodextrin. Images were taken at 25 °C. All the GUVs in the today are depleted of cholesterol, which is considered as a sort
figure have been labeled with the lipophilic fluorescence probes DiIC18 of “contaminant” for a good surface activity. Our results sug-
(red) and Bodipy-PC (green). C, fluorescence images of GUVs composed gest that last idea should be re-examined by using native
of a lipid fraction of NPSM that contains no surfactant proteins at material as model system.
different cholesterol molar ratio (with respect to phospholipids). Images
from left to right are from GUVs containing 10, 15, 18, 21, and 30 mol % The dependence of the lateral structure of surfactant mem-
cholesterol. Images were taken at 25 °C. The scale bars correspond to 10 branes on cholesterol concentration provides a framework for the
␮m. physiological meaning of rapid changes in the cholesterol-to-
phospholipid ratio reported previously (11) in pulmonary surfac-
TABLE I
Cholesterol content in NPSM and its fractions tant in response to various physiological stimuli. For example, in
homeothermic mammals, the cholesterol content changes with
% cholesterol/phospholipid
exercise, suggesting that changes in the mechanical properties of
molar ratioa surfactant membranes are required to accommodate the changes
NPSM 21.3 ⫾ 2 in ventilation. The cholesterol content in surfactant also varies
NPSM organic extract 19.2 ⫾ 3 with the temperature of the environment in both heterothermic
Protein-free NPSM organic extract 18.5 ⫾ 3
Cholesterol-depleted, protein-free NPSM 9.4 ⫾ 2
species (40) and in animals that enter torpor (41). The proportion
organic extract of cholesterol correlates with differences in lung structure, both
from a phylogenetic and an ontogenetic point of view (42). The
a
Values are means ⫾ S.D. after averaging four different samples of
each material. amount of cholesterol could be finely tuned in the context of the
presence, in given amounts, of other lipid species, such as satu-
rated and unsaturated phospholipids, to provide a structural
how better interfacial dynamics. Spreading kinetics were al- “scaffold” for surfactant membranes at precise conditions of tem-
ways faster at 37 than at 25 °C, but the difference in the ability perature and breathing dynamics. Potential physiological mech-
of the different samples to reach the highest surface pressures anisms providing tight regulatory control of surfactant composi-
persisted, suggesting that such effect is because of intrinsic tion, especially regarding cholesterol/phospholipid ratio, should
structural differences among the samples. Fig. 6C shows be investigated as they could be evolutionarily conserved to fit
spreading kinetics of NPSM suspensions spread on top of sub- surfactant material properties to required performance. For ex-
phases containing different concentrations of methyl-␤-cyclo- ample, compositional data reported for surfactant in the litera-
dextrin. Native surfactant spread rapidly at the first instance, ture include remarkable differences in terms of cholesterol-to
independent of the amount of methyl-␤-cyclodextrin present in phospholipid ratio, both among species and among individuals.
the subphase. However, the higher the amount of methyl-␤- The possibility that such differences could originate, at least
Phase Separation in Native Surfactant Membranes 40721

FIG. 6. Interfacial spreading ␲-t isotherms of NPSM and its fractions. ␲-t isotherms at 25 (A) and 37 °C (B) of NPSM (closed squares), its
reconstituted organic extract containing all the lipids plus the hydrophobic proteins (open squares), its reconstituted lipid fraction without proteins
(closed circles), and its reconstituted lipid fraction without proteins and partly depleted of cholesterol (10 mol % instead of 20%) (open circles). C,
␲-t isotherms of NPSM spread at 25 °C on top of subphases containing different concentrations of methyl-␤-cyclodextrin (indicated in the figure,
in mg/ml). All the experiments were done by spreading 3 ␮l of an aqueous suspension of the different materials (phospholipid concentration 10
mg/ml) on top of a 20-cm2 thermostated trough containing buffer (5 mM Tris, 150 mM NaCl, pH 7), with or without methyl-␤-cyclodextrin, as a
subphase. The subphase was equilibrated for 4 min at the desired temperature before initiating sample spreading.

partly, from environmental effects should be explored. These and hydrophobic proteins are involved in initiating monolayer-

Downloaded from http://www.jbc.org/ by guest on May 18, 2014


ideas need still to be fully established in future experiments, for to-bilayer transitions upon compression (during lung deflation)
example by comparing different surfactant preparations ex- and in forming structures that re-spread efficiently upon ex-
tracted from particular animal species at different environmen- pansion (during inspiration) (1, 46). Similarly, we speculate
tal conditions. that phase separation observed in surfactant membranes may
The thermotropic behavior of NPSM (Fig. 4) suggests that be relevant to provide particular regions (with particular com-
the composition of this material needs to be optimized to be at position and lateral arrangement) where accumulated SP-B
the “edge” of the liquid ordered/liquid disordered phase coex- and SP-C may initiate bilayer-to-monolayer molecular trans-
istence regime at physiological temperatures, which is likely to fer, a process that is essential in establishing the surface-active
provide particular structural and dynamical properties for its surfactant film in the lung. This last assumption is supported
mechanical function under particular environmental condi- by the connection between the results of the spreading exper-
tions. It is interesting to note that the phase transition event iments (Fig. 6) and the particular lateral structure of NPSM at
observed for NPSM is different from that observed for lipid different cholesterol-to-phospholipid molar ratios. We propose
mixtures displaying coexistence of two fluid phases (18, 21). In that in order to show optimal spreading behavior, the surfac-
the artificial system the fluid ordered/fluid disordered phase tant must contain hydrophobic surfactant proteins while si-
pattern (that is characterized for the presence of circular do- multaneously having a lipid composition permitting the coex-
mains) changes to a single homogeneous phase (the fluid dis- istence of fluid phases under physiological conditions. The role
ordered phase), an effect that is not observed in NPSM. The of the phase coexistence-containing structures in other surface-
possible maintenance of a minor proportion of ordered-like active or mechanical properties of surfactant still has to be
phase at temperatures a few degrees higher than 38 °C, such as explored in detail, but the present data indicate that composi-
those reached in feverish states, as well as the correlation tion of clinical surfactants designed to treat respiratory pathol-
between the structural features and the functional properties ogies such as acute respiratory distress syndrome could still be
of surfactant material remain to be established. For instance, largely improved.
Inoue et al. (43) showed that lung compliance was consistent Finally, our results also provide a clear example of the notion
with surfactant activity remaining intact above 42 °C. One that specialized regions with particular lateral packing prop-
could speculate that possible regulatory mechanisms might erties coexist in natural membranes and control specific molec-
exist to compensate for environmental factors through modifi- ular interactions in agreement with the raft hypothesis. The
cation of the relative proportion in surfactant of certain key raft hypothesis has been widely spread in recent years (13, 14)
species, with cholesterol a probable candidate. and suggests that factors modulating lipid segregation and
Cholesterol has been reported to induce a compression- domain dynamics may indirectly regulate specific membrane-
driven remixing behavior of coexisting phases in monolayers associated functions. There are remarkable aspects in this
composed of pulmonary surfactant organic extracts (8, 44). mechanism as we learned from our experiments with pulmo-
Such behavior is probably related to the existence of a cholesterol- nary surfactant membranes. First, the coexistence of defined
dependent critical miscibility pressure in the two-dimensional membrane regions with particular lateral arrangements is
phase diagrams of simple model systems (45). A possible cor- triggered by the presence of a few key lipid species (cholesterol,
relation between the compression-driven remixing and the DPPC instead of sphingomyelin, as in the raft hypothesis, and
temperature-induced melting of liquid ordered regions in sur- unsaturated phospholipids), independent of the compositional
factant bilayers observed in our experiments is difficult to complexity of the material. This last aspect could possibly be
attain because there is not a clear link between lateral pressure connected with events at supramolecular levels such as the
conditions in both systems (monolayers and bilayers). How- mechanical properties of the membrane. Second, the composi-
ever, both phenomena are indicative of surfactant composition tion of a particular membrane region displaying a particular
being, perhaps, evolutionarily tuned to support a highly dy- lateral structure (for example the liquid ordered or liquid dis-
namic behavior of the surfactant operative structures. ordered regions) would modulate specific molecular interac-
It has been proposed that the regions in pulmonary surfac- tions in the plane of the membrane (as seen in Fig. 5 with the
tant interfacial films accumulating unsaturated lipid species different distribution of the lipophilic probes between artificial
40722 Phase Separation in Native Surfactant Membranes
and native membranes displaying the coexistence of two fluid 3. Wright, J. R. (2003) J. Clin. Investig. 111, 1453–1455
4. Whitsett, J. A. & Weaver, T. E. (2002) N. Engl. J. Med. 347, 2141–2148
phases), possibly being connected with events occurring at the 5. Perez-Gil, J. & Keough, K. M. (1998) Biochim. Biophys. Acta 1408, 203–217
molecular level. Most interestingly, our results show that the 6. Lewis, J. F. & Veldhuizen, R. (2003) Annu. Rev. Physiol. 65, 613– 642
7. Discher, B. M., Maloney, K. M., Schief, W. R., Jr., Grainger, D. W., Vogel, V. &
surfactant proteins are not participating substantially in trig- Hall, S. B. (1996) Biophys. J. 71, 2583–2590
gering the particular lateral structure observed in this partic- 8. Nag, K., Perez-Gil, J., Ruano, M. L., Worthman, L. A., Stewart, J., Casals, C.
ular membrane. Even though this is a new observation in the & Keough, K. M. (1998) Biophys. J. 74, 2983–2995
9. Nag, K., Pao, J. S., Harbottle, R. R., Possmayer, F., Petersen, N. O. & Baga-
framework of the raft hypothesis, additional experiments in tolli, L. A. (2002) Biophys. J. 82, 2041–2051
other membrane systems are required to generalize this sug- 10. Piknova, B., Schram, V. & Hall, S. B. (2002) Curr. Opin. Struct. Biol. 12,
487– 494
gestion further. 11. Orgeig, S. & Daniels, C. B. (2001) Comp. Biochem. Physiol. A Mol. Integr.
CONCLUSIONS Physiol. 129, 75– 89
12. Casals, C., Arias-Diaz, J., Valino, F., Saenz, A., Garcia, C., Balibrea, J. L. &
We have obtained unambiguous evidence that at physiolog- Vara, E. (2003) Am. J. Physiol. 284, L466 –L472
13. Edidin, M. (2003) Annu. Rev. Biophys. Biomol. Struct. 32, 257–283
ical temperatures the lateral phase separation observed in 14. Simons, K. & Ikonen, E. (1997) Nature 387, 569 –572
native pulmonary surfactant membranes is entirely dependent 15. Korlach, J., Schwille, P., Webb, W. W. & Feigenson, G. W. (1999) Proc. Natl.
on the concentration of a few key lipid species, with a partic- Acad. Sci. U. S. A. 96, 8461– 8466
16. Bagatolli, L. A. & Gratton, E. (2000) Biophys. J. 78, 290 –305
ular involvement of cholesterol. Remarkably, this phenomenon 17. Bagatolli, L. A. & Gratton, E. (2000) Biophys. J. 79, 434 – 447
is independent of the presence of surfactant proteins. Choles- 18. Dietrich, C., Bagatolli, L. A., Volovyk, Z. N., Thompson, N. L., Levi, M.,
Jacobson, K. & Gratton, E. (2001) Biophys. J. 80, 1417–1428
terol extraction affects the spreading properties of the native 19. Feigenson, G. W. & Buboltz, J. T. (2001) Biophys. J. 80, 2775–2788
material suggesting an interesting correlation between the lat- 20. Veatch, S. L. & Keller, S. L. (2002) Phys. Rev. Lett. 89, 268101(-1)–268101(-4)
eral structure and one of the relevant physiological functions of 21. Veatch, S. L. & Keller, S. L. (2003) Biophys. J. 85, 3074 –3083
22. Scherfeld, D., Kahya, N. & Schwille, P. (2003) Biophys. J. 85, 3758 –3768
surfactant. Additionally, the dramatic lateral structure transi- 23. Baumgart, T., Hess, S. T. & Webb, W. W. (2003) Nature 425, 821– 824
tion observed at 37.5 °C underlines the critical properties as- 24. Casals, C., Herrera, L., Miguel, E., Garcia-Barreno, P. & Municio, A. M. (1989)
Biochim. Biophys. Acta 1003, 201–203
sociated with surfactant full composition under physiological 25. Rouser, G., Fkeischer, S. & Yamamoto, A. (1970) Lipids 5, 494 – 496

Downloaded from http://www.jbc.org/ by guest on May 18, 2014


conditions. Apparently, this material has been evolutionarily 26. Plasencia, I., Cruz, A., Lopez-Lacomba, J. L., Casals, C. & Perez-Gil, J. (2001)
optimized to sustain a structure that is at the borderline of a Anal. Biochem. 296, 49 –56
27. Düzgünes, N., Bagatolli, L. A., Meers, P., Oh, Y. & Straubinger, R. M. (2003)
lateral structure transition, likely contributing to essential lev- in Liposomes (Torchiling, V. P. & Weissig, V., eds) Vol. 2, pp. 105–147,
els of mechanical plasticity. We believe that our results point to Oxford University Press, London
28. Angelova, M. I. & Dimitrov, D. S. (1986) Faraday Discuss. Chem. Soc. 81,
an entirely new perspective on the establishment of structure- 303–311
function relationships in pulmonary surfactant, which may 29. Angelova, M. I., Soléau, S., Meléard, P. H., Faucon, J. F. & Bothorel, P. (1992)
significantly facilitate the search for new pulmonary surfactant Colloid Polym. Sci. 89, 127–131
30. Discher, B. M., Maloney, K. M., Grainger, D. W., Sousa, C. A. & Hall, S. B.
materials effective in the treatment of pathologies such as (1999) Biochemistry 38, 374 –383
acute respiratory distress syndrome. In addition, the relevance 31. Perez-Gil, J., Cruz, A. & Casals, C. (1993) Biochim. Biophys. Acta 1168,
261–270
of our results goes beyond the pulmonary surfactant field, into 32. Schurch, S., Green, F. H. & Bachofen, H. (1998) Biochim. Biophys. Acta 1408,
the general field of biomembranes. According to our results, 180 –202
pulmonary surfactant could be one of the first membranous 33. Nag, K., Taneva, S. G., Perez-Gil, J., Cruz, A. & Keough, K. M. (1997) Biophys.
J. 72, 2638 –2650
systems reported where the coexistence of specialized mem- 34. Morrow, I. C., Rea, S., Martin, S., Prior, I. A., Prohaska, R., Hancock, J. F.,
brane domains exists as a structural basis for its function. James, D. E. & Parton, R. G. (2002) J. Biol. Chem. 277, 48834 – 48841
35. Trauble, H., Eibl, H. & Sawada H. (1974) Naturwissenschaften 61, 344 –354
Acknowledgments—We thank Dr. I. Plasencia for technical assist- 36. Yu, S. H. & Possmayer, F. (2003) J. Lipid Res. 44, 621– 629
ance with protein and lipid analysis of surfactant fractions and Dr. Lars 37. Diemel, R. V., Snel, M. M., Van Golde, L. M., Putz, G., Haagsman, H. P. &
Batenburg, J. J. (2002) Biochemistry 41, 15007–15016
Duelund for the assistance with differential scanning calorimetry ex- 38. Taneva, S. & Keough, K. M. (1997) Biochemistry 36, 912–922
periments. We also thank Dr. David Jameson (University of Hawaii, 39. Yu, S. H. & Possmayer, F. (2001) J. Lipid Res. 42, 1421–1429
Manoa), Dr. Kevin M. W. Keough (Memorial University of Newfound- 40. Daniels, C. B., Barr, H. A., Power, J. H. & Nicholas, T. E. (1990) Exp. Lung Res.
land), Dr. Ole Mouritsen (University of Southern Denmark), Dr. Felix 16, 435– 449
Goñi (University of Basque Country), Dr. Enrico Gratton (University of 41. Langman, C., Orgeig, S. & Daniels, C. B. (1996) Am. J. Physiol. 271,
Illinois, Urbana-Champaign), and Drs. Sandra Orgeig and Chris R437–R445
Daniels (University of Adelaide) for their useful comments and sugges- 42. Orgeig, S., Daniels, C. B., Johnston, S. D. & Sullivan, L. C. (2003) Reprod.
Fertil. Dev. 15, 55–73
tions on a critical reading of the manuscript. MEMPHYS-Center for
43. Inoue, H., Inoue, C. & Hildebrandt, J. (1982) J. Appl. Physiol. 53, 567–575
Biomembrane Physics is supported by The Danish National Research 44. Discher, B. M., Schief, W. R., Vogel, V. & Hall, S. B. (1999) Biophys. J. 77,
Foundation. 2051–2061
45. McConnell, H. M. & Vrljic, M. (2003) Annu. Rev. Biophys. Biomol. Struct. 32,
REFERENCES 469 – 492
1. Perez-Gil, J. (2002) Biol. Neonate 81, Suppl. 1, 6 –15 46. Schief, W. R., Antia, M., Discher, B. M., Hall, S. B. & Vogel, V. (2003) Biophys.
2. Daniels, C. B. & Orgeig, S. (2003) News Physiol. Sci. 18, 151–157 J. 84, 3792–3806

You might also like