13.11.2015 Views

perez10_Surfactant

You also want an ePaper? Increase the reach of your titles

YUMPU automatically turns print PDFs into web optimized ePapers that Google loves.

Jesús Perez-Gil and Timothy E. Weaver<br />

Physiology 25:132-141, 2010. doi:10.1152/physiol.00006.2010<br />

You might find this additional information useful...<br />

This article cites 118 articles, 48 of which you can access free at:<br />

http://physiologyonline.physiology.org/cgi/content/full/25/3/132#BIBL<br />

This article has been cited by 1 other HighWire hosted article:<br />

Lamellar Bodies Form Solid Three-dimensional Films at the Respiratory Air-Liquid<br />

Interface<br />

A. Ravasio, B. Olmeda, C. Bertocchi, T. Haller and J. Perez-Gil<br />

J. Biol. Chem., September 3, 2010; 285 (36): 28174-28182.<br />

[Abstract] [Full Text] [PDF]<br />

Medline items on this article's topics can be found at http://highwire.stanford.edu/lists/artbytopic.dtl<br />

on the following topics:<br />

Biophysics .. Metabolism<br />

Medicine .. Pulmonary <strong>Surfactant</strong>s<br />

Physiology .. Homeostasis<br />

Physiology .. Pulmonary Alveoli<br />

Medicine .. Respiration<br />

Updated information and services including high-resolution figures, can be found at:<br />

http://physiologyonline.physiology.org/cgi/content/full/25/3/132<br />

Additional material and information about Physiology<br />

http://www.the-aps.org/publications/physiol<br />

This information is current as of September 9, 2010 .<br />

can be found at:<br />

Downloaded from physiologyonline.physiology.org on September 9, 2010<br />

Physiology (formerly published as News in Physiological Science) publishes brief review articles on major physiological<br />

developments. It is published bimonthly in February, April, June, August, October, and December by the American Physiological<br />

Society, 9650 Rockville Pike, Bethesda MD 20814-3991. Copyright © 2010 by the American Physiological Society. ISSN: 1548-9213,<br />

ESSN: 1548-9221. Visit our website at http://www.the-aps.org/.


REVIEWS<br />

Pulmonary <strong>Surfactant</strong> Pathophysiology:<br />

Current Models and Open Questions<br />

Pulmonary surfactant is an essential lipid-protein complex that stabilizes the<br />

respiratory units (alveoli) involved in gas exchange. Quantitative or qualitative<br />

derangements in surfactant are associated with severe respiratory pathologies.<br />

The integrated regulation of surfactant synthesis, secretion, and metabolism<br />

is critical for air breathing and, ultimately, survival. The goal of this<br />

review is to summarize our current understanding and highlight important<br />

knowledge gaps in surfactant homeostatic mechanisms.<br />

PHYSIOLOGY 25: 132–141, 2010; doi:10.1152/physiol.00006.2010<br />

Jesús Perez-Gil 1 and<br />

Timothy E. Weaver 2<br />

1 Department Bioquímica, Faculty Biología,<br />

Universidad Complutense, Madrid, Spain; and<br />

2 Cincinnati Children’s Research Foundation<br />

and University of Cincinnati College of<br />

Medicine, Cincinnati, Ohio<br />

jpg@bbm1.ucm.es<br />

The vital process of mammalian breathing is dependent<br />

on an extensive gas exchange surface provided<br />

by the alveoli in the lung periphery. Surface<br />

tension on the epithelial side of the air-blood barrier<br />

exerts a collapsing pressure that is stabilized by<br />

spreading of a lipid-rich film (pulmonary surfactant)<br />

at the alveolar air-liquid interface. Quantitative<br />

[e.g., respiratory distress syndrome (RDS)] or<br />

qualitative [e.g., acute RDS (ARDS)] changes in surfactant<br />

lead to alveolar collapse and the need for<br />

ventilatory support; likewise, accumulation of surfactant<br />

in the alveolar airspaces (e.g., pulmonary<br />

alveolar proteinosis) impedes gas exchange. Thus<br />

the integrated regulation of surfactant synthesis,<br />

secretion, and metabolism is essential for air<br />

breathing and, ultimately, survival.<br />

Research in the past three decades has provided<br />

extensive insight into surfactant biology that, in<br />

turn, has facilitated development of surfactant replacement<br />

therapies: the latter have profoundly<br />

impacted the incidence of morbidity and mortality<br />

in newborn infants. However, despite the many<br />

remarkable achievements in this field, major questions<br />

remain unanswered. The goal of this review is<br />

to summarize our present understanding of surfactant<br />

homeostatic mechanisms and highlight important<br />

knowledge gaps in molecular pathways<br />

involved in surfactant synthesis, secretion, recycling,<br />

and degradation. This focus excludes discussion<br />

of the role of surfactant in innate lung<br />

defense, and the reader is referred to several excellent<br />

reviews of this topic (46, 57, 60, 109).<br />

Synthesis and Assembly<br />

of <strong>Surfactant</strong><br />

The major components of pulmonary surfactant<br />

include phospholipids (80%), neutral lipids<br />

(mainly cholesterol, 10%), and the two hydrophobic<br />

peptides (1–2%) surfactant protein B (SP-B)<br />

and SP-C. Type II epithelial cells synthesize and<br />

assemble the lipid and protein components into<br />

complexes that are stored as tightly packed membranes<br />

in lamellar bodies until secreted into the<br />

alveolar airspaces (FIGURE 1). Molecular pathways<br />

involved in the regulation of expression, intracellular<br />

trafficking, and processing of the surfactant<br />

proteins have been extensively characterized;<br />

much less is known about the intracellular pathways<br />

regulating surfactant lipid biosynthesis.<br />

<strong>Surfactant</strong> Proteins<br />

<strong>Surfactant</strong> Protein B<br />

SP-B is synthesized as a soluble proprotein that is<br />

processed to a smaller, lipid-associated peptide in<br />

the distal secretory pathway within the type II cell.<br />

Processing of proSP-B occurs in the multivesicular<br />

body (MVB) and lamellar body (LB) compartments<br />

and involves the action of napsin A, cathepsin H,<br />

and pepsinogen C (12, 37, 45, 96); at least one and<br />

perhaps several other as yet unknown enzymes<br />

participate in proprotein maturation. The fully<br />

processed, 79-amino acid mature peptide facilitates<br />

organization of surfactant membranes in the<br />

lamellar body, likely through its ability to promote<br />

membrane-membrane contacts, perturbation of<br />

lipid packing, and membrane fusion (49, 89). SP-B<br />

deficiency, arising from mutations in the human<br />

gene (SFTPB) or disruption of the mouse locus<br />

(Sftpb), results in vesiculated lamellar bodies with<br />

few or no bilayer membranes and electron dense<br />

inclusions. Most importantly, severe SP-B deficiency<br />

(i.e., 75% reduction in SP-B content in the<br />

airspaces) results in fatal RDS (19, 67, 70, 71). SP-B<br />

expression is altered in a number of acute and<br />

chronic lung diseases and may therefore contribute<br />

to pathogenesis (107). Thus identification of<br />

transcriptional pathways that modulate SP-B expression<br />

and proprotein maturation in both<br />

healthy and diseased lung remains an important<br />

area of investigation.<br />

SP-B is also expressed in non-ciliated, bronchiolar<br />

epithelial cells (Clara cells). The proprotein is<br />

Downloaded from physiologyonline.physiology.org on September 9, 2010<br />

132<br />

1548-9213/10 ©2010 Int. Union Physiol. Sci./Am. Physiol. Soc.


REVIEWS<br />

not fully processed to the mature peptide in Clara<br />

cells, and the processing enzymes and fates of resulting<br />

peptides are largely unknown. Importantly,<br />

targeted expression of SP-B in type II epithelial<br />

cells but not Clara cells of Sftpb / mice is sufficient<br />

to rescue the neonatal lethal phenotype (62).<br />

Taken together, these data suggest that the function<br />

of SP-B in Clara cells is unrelated to surfactant<br />

homeostasis. Given the presence of saposin-like<br />

domains in the NH 2 - and COOH-terminal regions<br />

of the proprotein (77), it is possible that saposinlike<br />

peptides derived from these regions may play a<br />

role in innate host defense of the airspaces. Consistent<br />

with this hypothesis, a saposin-like peptide<br />

derived from the SP-B propetide was recently identified<br />

in bronchoalveolar lavage fluid and alveolar<br />

macrophages. The peptide exhibited antimicrobial<br />

activity only at acidic pH, suggesting that it promotes<br />

killing of bacteria internalized by macrophages in the<br />

airspaces (112). Whether the saposin-like domain in<br />

the COOH-terminal region of the SP-B proprotein is<br />

also involved in alveolar defense remains to be<br />

established.<br />

<strong>Surfactant</strong> Protein C<br />

Unlike SP-B, SP-C expression in the lung is restricted<br />

to alveolar type II epithelial cells. Newly<br />

synthesized SP-C is an integral membrane protein<br />

in which the NH 2 terminus resides in the cytosol<br />

and the COOH terminus resides in the ER lumen<br />

(56). The cytosolic domain encodes information<br />

required for intracellular trafficking of SP-C to the<br />

MVB/LB compartments (20, 21, 54, 58). The COOH<br />

terminal, lumenal domain may act as an intramolecular<br />

chaperone that transiently stabilizes the inherently<br />

unstable transmembrane domain (TMD),<br />

presumably until palmitoylation of two cytosolic<br />

cysteine residues assume this role (44, 52, 53). The<br />

identity of the palmitoyl acyltransferase that acylates<br />

SP-C is not known (93).<br />

Like SP-B, SP-C is synthesized as a proprotein<br />

that is processed to a smaller, hydrophobic peptide<br />

in the MVB/LB compartments. The fully processed,<br />

35-amino acid mature peptide consists of a TMD<br />

and a short segment of the cytosolic domain. Enzymes<br />

involved in the maturation of proSP-C have<br />

Downloaded from physiologyonline.physiology.org on September 9, 2010<br />

FIGURE 1. Biosynthesis of pulmonary surfactant<br />

Synthesis of the protein and phospholipid components of surfactant may proceed via separate pathways (green<br />

lines). SP-B and SP-C traffic through the Golgi and late endosome/multivesicular body (MVB) to the lamellar body. In<br />

contrast, surfactant phospholipids may traffic directly from the ER to the lamellar body; phospholipid transfer protein(s)<br />

(PLTP) and ABC transporters likely play an important role in phospholipid trafficking. (It is also possible that<br />

direct contact between the ER and lamellar body may facilitate phospholipid transfer; see text.) Transcriptional pathways<br />

involved in coordinated regulation of surfactant protein and phospholipid synthesis are indicated by black arrows.<br />

<strong>Surfactant</strong> components internalized from the airspaces may also contribute to biosynthesis (red lines). The<br />

micrograph of the multilayered surfactant surface film was modified from Ref. 88.<br />

PHYSIOLOGY • Volume 25 • June 2010 • www.physiologyonline.org 133


REVIEWS<br />

not been conclusively identified, although cathepsin<br />

H has been implicated (54), suggesting that the<br />

processing machinery may be shared with SP-B.<br />

The membrane localization of SP-C presents an<br />

interesting obstacle to its secretion. This conundrum<br />

is solved by internalization of the proprotein<br />

from the limiting membrane of the MVB to small<br />

lumenal vesicles, an event that is regulated by<br />

Nedd4-2-mediated mono-ubiquitination of proSP-C<br />

(20, 58). Fusion of the MVB with a LB leads to SP-Bmediated<br />

incorporation of SP-C-containing lumenal<br />

vesicles into existing surfactant membranes that are<br />

eventually secreted into the airspaces. Thus mature<br />

SP-C, one of the most hydrophobic peptides known,<br />

maintains its transmembrane orientation throughout<br />

biosynthesis.<br />

Surprisingly, disruption of the SP-C locus in<br />

mice (Sftpc) has no overt effect on synthesis, trafficking,<br />

packaging, or secretion of surfactant by<br />

type II epithelial cells (40, 41). Sftpc / mice survive<br />

with essentially normal gas exchange apparently<br />

because SP-B is sufficient for surfactant<br />

function in vivo. However, SP-C deficiency is associated<br />

with development of chronic lung disease in<br />

humans (3) and strain-dependent, progressive lung<br />

disease in Sftpc / mice. Identification of genes that<br />

confer strain-dependent sensitivity and/or resistance<br />

to lung disease in Sftpc / mice remains an important<br />

and unresolved issue (44). <strong>Surfactant</strong> films from<br />

SP-C-deficient mice show minor intrinsic instability<br />

(40), which may contribute to chronic respiratory<br />

disease; however, it is equally possible that SP-C has<br />

an unidentified surfactant-independent function(s).<br />

Mutations in the SFTPC gene have been linked to<br />

chronic lung disease, possibly through formation of<br />

cytotoxic aggregates of mutant SP-C proprotein (for<br />

review, see Ref. 107). Overall, conservation of SP-C<br />

among species is consistent with an important function;<br />

however, despite considerable information<br />

about SP-C biosynthesis and in vitro surface activity,<br />

the role of this peptide in surfactant homeostasis and<br />

lung disease remains obscure.<br />

Synthesis of <strong>Surfactant</strong> Lipids<br />

Phosphatidylcholine<br />

Phosphatidylcholine (PC) is a major component of<br />

all cellular membranes (98) and is the most abundant<br />

surfactant phospholipid (100). The rate limiting<br />

enzyme for PC synthesis, cholinephosphate<br />

cytidylyltransferase, is expressed as two isoforms<br />

(CCT and CCT) encoded by separate genes. Targeted<br />

disruption of the CCT locus (Pcytla) orthe<br />

locus encoding another key enzyme in this pathway,<br />

choline kinase (Chka), results in embryonic<br />

lethality confirming that PC synthesis is essential<br />

for early embryogenesis (104, 110). Targeted deletion<br />

of Pcytla in type II epithelial cells of mice<br />

resulted in neonatal RDS and death within 10 min<br />

of birth (94). Lung development in Pcytla / mice<br />

was normal, but total PC content in the airspaces<br />

was dramatically reduced and formation of lamellar<br />

bodies in type II epithelial cells was highly aberrant.<br />

These results confirm the key role of the<br />

CCT isoform in synthesis of surfactant PC.<br />

Approximately half of the PC in surfactant is<br />

dipalmitoylphosphatidylcholine (DPPC; PC 16:0/<br />

16:0), accounting for 40% of total surfactant<br />

phospholipid (82, 100). DPPC is the lipid considered<br />

as primarily responsible for the surface tension<br />

reducing properties of surfactant; therefore,<br />

regulation of this PC species is likely critical for<br />

surfactant homeostasis. Two pathways contribute<br />

to the production of DPPC, de novo synthesis and<br />

remodeling via lysoPC. The latter pathway accounts<br />

for up to 75% of DPPC in type II cells and<br />

involves deacylation of unsaturated PC at the sn-2<br />

position by a calcium-dependent phospholipase<br />

A2 (33); peroxiredoxin 6, a calcium-independent<br />

phospholipase A2, may also contribute to deacylation<br />

of PC (25, 34). The enzyme responsible for<br />

re-acylation, lysophosphatidylcholine acyltransferase,<br />

was recently identified and named LPCATI<br />

(69). This finding is a significant achievement that<br />

will facilitate testing of a number of hypotheses in<br />

vivo. Targeted disruption of the LPCATI locus in<br />

type II epithelial cells will permit assessment of the<br />

importance of the remodeling pathway in surfactant<br />

homeostasis, identification of potential compensatory<br />

pathways, and the overall importance of<br />

DPPC for surfactant function in vivo. In addition,<br />

recent evidence supports the hypothesis that there<br />

is cross-talk between the remodeling and the de<br />

novo synthesis pathways (16). Identification of<br />

transcriptional networks that integrate expression<br />

of LPCATI and other lipogenic enzymes with surfactant<br />

proteins should provide important insight<br />

into the molecular pathways governing surfactant<br />

homeostasis (FIGURE 1).<br />

Phosphatidylglycerol<br />

The second most abundant surfactant phospholipid,<br />

PG, comprises 10% of the surfactant lipid<br />

fraction in humans. This anionic phospholipid is<br />

not a common component of intracellular membranes<br />

(98). Unlike PC, surfactant PG content and<br />

molecular species varies significantly among animals<br />

(82). However, the overall content of anionic phospholipid<br />

(PGPI) is relatively constant, suggesting<br />

that the contribution of acidic polyhydroxylated<br />

phospholipids is more important than the precise<br />

molecular forms. A key step in PG synthesis is<br />

conversion of CDP-diacylglycerol to phosphatidylglycerolphosphate<br />

by phosphatidylglycerophosphate<br />

synthase (8, 9). Transcription of the gene<br />

encoding this enzyme (PGS1) is increased threefold<br />

Downloaded from physiologyonline.physiology.org on September 9, 2010<br />

134<br />

PHYSIOLOGY • Volume 25 • June 2010 • www.physiologyonline.org


REVIEWS<br />

during maturation of human lung epithelial cells in<br />

vitro (103) and is similarly increased in rat type II<br />

cells when cultured under conditions that promote<br />

differentiation (65). Targeted disruption of the Pgs1<br />

locus in type II epithelial cells could provide important<br />

new information regarding the role of PG<br />

in LB maturation, surfactant function, and the potential<br />

for PI to compensate for loss of PG. Although<br />

the role of PG in surfactant homeostasis<br />

remains unclear, two recent studies presented<br />

compelling preliminary evidence that PG can suppress<br />

viral infection and inflammatory responses<br />

in the lung (61, 72). These novel findings underscore<br />

the importance of filling this particular<br />

knowledge gap.<br />

Intracellular Trafficking of<br />

<strong>Surfactant</strong> Phospholipids<br />

The fully assembled surfactant lipid-protein complex<br />

is stored as tightly packed bilayer membranes<br />

in lamellar bodies of type II epithelial cells. Since<br />

lamellar bodies lack the requisite enzymes for<br />

phospholipid biosynthesis (8), surfactant lipids<br />

must be transported from the site of synthesis in<br />

the ER to the storage organelle. The identity of<br />

pathways involved in intracellular trafficking of<br />

surfactant phospholipids remains one of the major<br />

unanswered questions in this field. Transport of<br />

intracellular lipids can occur by at least three pathways,<br />

including vesicular transport, non-vesicular<br />

transport, and diffusion at membrane contact sites (98).<br />

Vesicular transport clearly plays an important<br />

role in trafficking of newly synthesized surfactant<br />

proteins via the Golgi to the lamellar body (FIGURE 1).<br />

Vesicular transport is also important for recycling<br />

of surfactant components (both proteins and lipids)<br />

from the alveolar airspaces to the LB via the<br />

endocytic pathway. The biosynthetic and endocytic<br />

pathways converge at the MVB/late endosome<br />

(the boundary between these endosomal<br />

compartments is not well defined in type II epithelial<br />

cells). The internal membranes of MVBs are<br />

enriched in cholesterol and lyso-bisphosphatidic<br />

acid. Thus the endocytic recycling pathway is likely<br />

an important source of cholesterol and minor<br />

phospholipid components of lamellar body membranes<br />

(FIGURE 1).<br />

In contrast to vesicular transport of surfactant<br />

lipids in the endocytic pathway, there is very little<br />

evidence for anterograde vesicular transport of<br />

DPPC or PG from the ER to the LB. Ultrastructural,<br />

autoradiographic analyses of type II epithelial cells<br />

in vivo detected the appearance of 3 [H]cholinelabeled<br />

lipids first in the ER followed by the Golgi<br />

and LB; notably label was not detected in the MVB<br />

(27). Although these results are consistent with<br />

trafficking of PC directly to the LB and/or via the<br />

Golgi, a separate study found that inhibition of<br />

vesicular transport through the Golgi did not inhibit<br />

incorporation of newly synthesized lipid into<br />

the LB (75). A unifying hypothesis posits that PC<br />

moves directly from the ER to the LB via a nonvesicular<br />

transport pathway. This hypothesis is supported<br />

by the localization of the ABC transporter,<br />

ABCA3, to the limiting membrane of the LB (68,<br />

111). Disruption of the Abca3 locus results in neonatal<br />

lethal RDS associated with loss of mature<br />

lamellar bodies and a dramatic decrease in production<br />

of lung PG and PC (7, 26, 35, 48). Similar<br />

findings have been reported for humans with severe,<br />

recessive ABCA3 mutations (13–15, 32, 39, 66,<br />

87, 91). Since ABC transporters typically move substrate<br />

out of the cytosol, it is highly likely that<br />

ABCA3 is an essential transporter of PC and/or PG<br />

into the LB.<br />

How is PC and/or PG transported from the ER to<br />

ABCA3 in the LB membrane? Intermembrane<br />

transport of PC and PG is significantly elevated in<br />

type II epithelial cells relative to whole lung and<br />

transfer activity increases commensurate with surfactant<br />

synthesis before birth (38, 63). Intracellular<br />

phospholipid transport activity was generally attributed<br />

to phospholipid transfer proteins and to<br />

PC-TP/StarD2 in particular. However, lung structure<br />

and function, including LB morphology and<br />

alveolar DPPC content, was unaffected in Pc-tp /<br />

mice (97). This outcome was quite unexpected,<br />

given the specificity of PC-TP/StarD2 for PC (55)<br />

and its tempo-spatial relationship to surfactant<br />

synthesis. It is currently not clear whether compensation<br />

occurs in type II cells of Pc-tp / mice<br />

or whether PC transfer is normally accomplished<br />

independently of PC-TP/StarD2. Other members<br />

of the START domain family (2) could account for<br />

PC transfer activity in type II epithelial cells, in<br />

particular StarD10. It is also possible that this function<br />

could be fulfilled by phospholipid transfer<br />

proteins lacking a START domain. SCP-2/nsL-TP<br />

and PI-TP both have PC transfer activity (108), and<br />

the former is enriched in type II epithelial cells and<br />

increases during fetal lung maturation (10). Overall,<br />

the involvement of phospholipid transfer proteins<br />

in nonvesicular transport of newly synthesized<br />

phospholipids from the ER to the LB remains an<br />

important and unresolved question.<br />

Another possible mechanism for transfer of<br />

phospholipids between donor and acceptor membranes<br />

is via diffusion or facilitated exchange at<br />

membrane contact sites (98) (FIGURE 1). This<br />

transfer mechanism would likely require the participation<br />

of accessory proteins to ensure the formation<br />

of transient contact sites between the<br />

appropriate donor (ER) and acceptor (LB) organelles.<br />

It has been suggested that some phospholipid<br />

transfer proteins could fulfill the role of<br />

Downloaded from physiologyonline.physiology.org on September 9, 2010<br />

PHYSIOLOGY • Volume 25 • June 2010 • www.physiologyonline.org 135


REVIEWS<br />

both accessory and transfer protein (2). Regardless<br />

of whether membrane contact is necessary for targeted<br />

transport of PC or PG from the ER to the LB,<br />

it is likely that a transfer protein(s) with docking<br />

motifs specific for each organelle would be required.<br />

Lamellar Body Maturation<br />

Based on the foregoing discussion, a simple model<br />

of lamellar body maturation can be constructed<br />

(FIGURE 1). The major surfactant phospholipids<br />

(DPPC, unsaturated PC, and PG) are synthesized in<br />

the ER. Because there is very little DPPC or PG in<br />

the ER, newly synthesized phospholipid must be<br />

rapidly exported from this organelle. Phospholipid<br />

transport is likely facilitated by specific phospholipid<br />

transfer proteins containing docking motifs<br />

that enable cargo loading at the ER and cargo<br />

discharge at the LB. Although phospholipid transfer<br />

proteins typically carry only a single cargo molecule,<br />

transfer can occur very rapidly (5). Import of<br />

PC and PG into the LB likely occurs through<br />

ABCA3. An excess of phospholipids at the inner<br />

leaflet of the LB membrane could drive the assembly<br />

of internal surfactant membranes (78). Incorporation<br />

of cholesterol and minor phospholipid<br />

components into surfactant membranes could be<br />

mediated by ABCA3, by unidentified lipid transporters<br />

in the LB membrane, or via the endocytic/<br />

recycling pathway. The process leading to highly<br />

organized, tightly packed surfactant membranes is<br />

poorly understood but likely requires SP-B and the<br />

fluidizing effects of cholesterol/minor phospholipids.<br />

This highly simplified model raises numerous<br />

questions. Do newly synthesized minor surfactant<br />

lipid components employ the same transport machinery<br />

(i.e., phospholipid transfer proteins and<br />

ABCA3) as PC and PG? Are PC and PG carried by<br />

the same phospholipid transfer protein or are multiple<br />

carrier proteins involved? Is phospholipid import<br />

into the LB accomplished through more than<br />

one ABC transporter? Are elevated calcium levels in<br />

the LB required for packing of negatively charged<br />

(PG-rich) membranes and, if so, what is the molecular<br />

identity of the calcium pump? How is LB<br />

size (maturation) monitored/regulated? Proteomic<br />

analyses of the limiting membrane of the LB, currently<br />

underway in several laboratories (105),<br />

should identify candidate proteins that, in turn, may<br />

provide answers to these and related questions.<br />

Extracellular <strong>Surfactant</strong><br />

Reorganization of <strong>Surfactant</strong> Membranes<br />

in the Alveolar Airspaces<br />

Transition of the intracellular storage form of surfactant<br />

(bilayer membranes in LBs) to an extracellular<br />

functional surface film has been extensively<br />

studied but is only partly understood (78, 114). The<br />

alveolar epithelial cell surface is covered by a thin<br />

aqueous layer (hypophase) in which newly secreted,<br />

tightly packed surfactant membranes reorganize<br />

to form a loose network of interconnected<br />

membranes that contacts the air-liquid interface. It<br />

is not clear how the structural transformation of<br />

surfactant membranes is triggered, but potential<br />

effectors include changes in the hydration of surfactant<br />

complexes, pH, or calcium concentration<br />

(29–31, 78, 84, 85). Subsequently, SP-B and SP-C<br />

are thought to facilitate transfer of surface active<br />

molecules (particularly phospholipids and, most<br />

importantly, DPPC) from the membrane network<br />

to the surface film (see FIGURE 2) (79, 89). Although<br />

this model of interfacial adsorption is<br />

widely accepted, translation of the results of in<br />

vitro analyses to surfactant adsorption in vivo is<br />

not straightforward. In particular, most in vitro<br />

studies have used surfactant preparations that are<br />

significantly diluted (simulating unpacked surfactant<br />

membranes) and simplified (i.e., simple phospholipid<br />

mixtures) compared with newly secreted<br />

surfactant. Moreover, the cellular microenvironment<br />

(composition and volume of the hypophase)<br />

may profoundly influence transfer of phospholipids<br />

from the packed membrane state to the surface<br />

film (33). In this regard, the results of recent studies<br />

revealed that packed surfactant membranes<br />

(similar to newly secreted surfactant) can spontaneously<br />

adsorb and rapidly transfer surface active<br />

material into the air-liquid interface, apparently<br />

without passing through an unpacked state (29,<br />

47). Thus the mechanisms by which newly secreted<br />

surfactant phospholipids are transferred to the airliquid<br />

interface in vivo remains an open question.<br />

Composition and Function of the<br />

Surface-Active Film<br />

Cyclical expansion (inhalation) and contraction<br />

(exhalation) of alveoli lead to changes in alveolar<br />

surface area accompanied by altered surface tension.<br />

Early work by Clements (19a) proposed that<br />

surface tension must be 2 mN/m at low lung<br />

volumes to prevent alveolar collapse (atelectasis)<br />

and rise to a maximum of 20–25 mN/m at higher<br />

volumes to maintain alveolar stability. This goal is<br />

achieved by maintenance of an interfacial film<br />

highly enriched in DPPC, a saturated phospholipid<br />

that produces extremely low surface tension on<br />

film compression (1 mN/m). The lack of fluidity<br />

of pure DPPC films at body temperature is overcome<br />

by incorporation of phospholipids with unsaturated<br />

acyl chains and cholesterol into surfactant<br />

complexes and the interfacial film (10a, 27a, 30a,<br />

68a). Observation of surfactant films by epifluorescence<br />

or atomic force microscopy suggests that satu-<br />

Downloaded from physiologyonline.physiology.org on September 9, 2010<br />

136<br />

PHYSIOLOGY • Volume 25 • June 2010 • www.physiologyonline.org


REVIEWS<br />

rated phospholipids (DPPC) segregate into ordered<br />

micro- and nano-domains, which can be highly<br />

compressed to produce the requisite low surface tension;<br />

unsaturated phospholipids and surfactant proteins<br />

remain in disordered regions that facilitate<br />

rapid transfer of phospholipids into (and out of) the<br />

surface film (22a, 27a, 115). Comingling of ordered<br />

and disordered domains thus provides a<br />

potential mechanism for maintenance of a stable<br />

interfacial film.<br />

Although we have a basic understanding of surfactant<br />

film dynamics, many details of the underlying<br />

processes remain unclear. It is assumed that<br />

DPPC enrichment of the surface film is required to<br />

support maximal pressures (very low surface tensions)<br />

at end expiration, but the actual proportion<br />

of DPPC that reaches the interfacial film is not<br />

known. Some studies suggested that SP-B and<br />

SP-C could selectively promote insertion of DPPC<br />

(99) into the interfacial film, but direct proof for<br />

this activity is lacking. It has been also proposed<br />

that compression of the surface film during expiration<br />

induces progressive refining of the interfacial<br />

film by squeezing out less stable unsaturated<br />

species, leading to DPPC enrichment (28, 64, 76).<br />

Direct examination of the structure and composition<br />

of the surface film formed from native surfactant<br />

under physiologically relevant conditions is<br />

required to confirm or discard these hypotheses. It<br />

is currently thought that the ability of surfactant to<br />

produce very low tensions during breathing is dependent<br />

on formation of a multilayered film at the<br />

interface (6, 36, 88). However, a comparison of the<br />

rheological properties of multilayered vs. monolayered<br />

surfactant films is required to test this model.<br />

Finally, an active line of research has tried to assign<br />

specific activities to SP-B and SP-C in remodeling<br />

of interfacial surfactant films during compression<br />

and facilitating film re-spreading during expansion<br />

(see FIGURE 3). Consistent with the neonatal lethal<br />

phenotype of SP-B-deficient mice, in vitro approaches<br />

using devices that capture some of the<br />

physiological compression-expansion constraints<br />

of the respiratory interface revealed that SP-B is<br />

required to achieve and sustain the lowest tensions<br />

during compression. SP-C, on the other hand, may<br />

cooperate in facilitating interfacial surfactant dynamics,<br />

including folding of the surface film, maintaining<br />

the multilayer reservoir attached to the<br />

interface during compression and promoting efficient<br />

re-spreading of the film on expansion. Optimal<br />

performance of surfactant at the interface may<br />

therefore require cooperation of both hydrophobic<br />

surfactant proteins (1). Validation of these models in<br />

vivo requires the development of new genetically<br />

modified animal models that are not yet available.<br />

<strong>Surfactant</strong> Recycling and <strong>Surfactant</strong><br />

Homeostasis<br />

Compression-expansion cycling leads to progressive<br />

conversion of the surface active fractions of<br />

Downloaded from physiologyonline.physiology.org on September 9, 2010<br />

FIGURE 2. Pulmonary surfactant adsorption to the interface and surface film formation<br />

Processes that may contribute to transport of surface active surfactant species to the interface include 1) direct cooperative<br />

transfer of surfactant from secreted lamellar body-like particles touching the interface, 2) unraveling of secreted<br />

lamellar bodies to form intermediate structures such as tubular myelin (TM) or large surfactant layers that have<br />

the potential to move and transfer large amounts of material to the interface, and 3) rapid movement of surface active<br />

species through a continuous network of surfactant membranes, a so-called surface phase, connecting secreting<br />

cells with the interface.<br />

PHYSIOLOGY • Volume 25 • June 2010 • www.physiologyonline.org 137


REVIEWS<br />

surfactant into much less active lipid/protein<br />

complexes (43, 95). Inactivation probably includes<br />

detachment of small particulate entities from the<br />

interface, changes in lipid/protein organization,<br />

and oxidation of lipid and protein species on repetitive<br />

exposure to air (83). <strong>Surfactant</strong> may also be<br />

inactivated by incorporation of materials inhaled<br />

from the upper airways or leaked from capillaries<br />

(42, 114). Maintenance of a fully functional surfactant<br />

film therefore requires continual film refinement<br />

through efficient removal of spent surfactant<br />

and incorporation of newly secreted complexes. It<br />

remains unclear whether the distinct surfactant<br />

structures that coexist in the alveolar pool are specifically<br />

targeted to functional, recycling, or clearance<br />

pathways.<br />

<strong>Surfactant</strong> homeostasis requires coordinated<br />

regulation of the processes summarized in<br />

FIGURE 1. The alveolar surfactant pool is continuously<br />

depleted through cellular uptake by type<br />

II epithelial cells and alveolar macrophages as<br />

well as removal via the mucociliary escalator. It is<br />

unclear whether cellular uptake of alveolar surfactant<br />

occurs indiscrimately or whether there is selective<br />

targeting of specific forms of surfactant to<br />

type II cells for recycling and/or macrophages for<br />

degradation. Preliminary support for selective uptake<br />

of alveolar surfactant by type II cells comes<br />

from recent studies in SP-D-deficient mice (50, 51).<br />

Interaction of the collectin SP-D with PI-rich surfactant<br />

complexes altered surfactant structure and<br />

enhanced uptake by type II cells but not macrophages.<br />

Thus, in addition to its well documented<br />

role in host defense, SP-D plays an important role<br />

in regulating alveolar pool size.<br />

The loss of surfactant from the airspaces is balanced<br />

by secretion of surfactant stored in lamellar<br />

bodies: regulation of surfactant secretion is therefore<br />

critical for homeostasis. Several excellent, recent<br />

reviews provide a comprehensive analysis of<br />

surfactant secretion and the central role of calcium<br />

mobilization in this process (4, 29). However, although<br />

pharmacological and physical stimuli of surfactant<br />

secretion have been extensively characterized,<br />

Downloaded from physiologyonline.physiology.org on September 9, 2010<br />

FIGURE 3. Participation of proteins SP-B and SP-C in surfactant dynamics during respiratory compression-expansion<br />

cycling<br />

Hydrophobic surfactant proteins play key roles in facilitating optimal dynamic behavior of surfactant during breathing.<br />

During exhalation, the surfactant surface film folds through three-dimensional transitions, forming a complex structure<br />

that sustains maximal pressures. Protein SP-C facilitates compression-driven folding of surfactant interfacial films (1),<br />

expulsion of lipids and lipid/protein complexes from the interface (80, 92), and formation of three-dimensional phases<br />

(59, 101, 102). SP-B stabilizes the interfacial film (2) and promotes establishment of membrane-membrane contacts<br />

(22, 24, 86), leading to formation of multilayered membrane arrays (3) (17). SP-B seems to provide mechanical stability<br />

to compressed films (23), possibly through increasing cohesivity between surfactant layers during maximal compression<br />

of the surface film. SP-C could promote association of excluded surfactant structures with the maximally<br />

compressed interface, likely through its NH 2 -terminal segment and/or insertion of palmitoylated cysteines of the protein<br />

into tightly packed interfacial films (4) (11, 102). Finally, both SP-B (23, 90, 106) and SP-C (73, 74, 81) seem to<br />

promote insertion (5) and re-spreading (6) of phospholipids from subsurface compartments into the interface during<br />

expansion.<br />

138<br />

PHYSIOLOGY • Volume 25 • June 2010 • www.physiologyonline.org


REVIEWS<br />

the molecular pathway(s) linking changes in alveolar<br />

pool size to stimulation of secretion has not been<br />

identified.<br />

Depletion of the intracellular pool of surfactant<br />

via the secretory pathway is balanced in part by<br />

uptake and recycling of alveolar surfactant in type<br />

II cells. Recycling is not sufficient to replenish the<br />

intracellular pool because some internalized surfactant<br />

is degraded: how internalized surfactant is<br />

partitioned between the recycling and degradation<br />

pathways remains an important unresolved question.<br />

To maintain intracellular pool size, new synthesis<br />

of surfactant must be coupled to recycling.<br />

<strong>Surfactant</strong> pool size also varies significantly with<br />

age, the alveolar pool being much larger in neonates<br />

than in adults (18, 50, 113). Overall, maintenance<br />

of the balance between the intracellular<br />

storage pool and the functional alveolar pool requires<br />

tight integration of synthesis, secretion, recycling,<br />

and degradation. How these pathways are<br />

modulated in response to quantitative and/or<br />

qualitative changes in alveolar surfactant remains<br />

a large and important knowledge gap.<br />

In summary, although much has been learned<br />

regarding the composition and function of surfactant,<br />

virtually nothing is known about the transcriptional<br />

networks that integrate molecular<br />

pathways involved in prenatal type II cell maturation,<br />

surfactant protein and lipid synthesis, and<br />

alveolar defense. How these molecular networks<br />

“sense” alveolar pool size in the postnatal lung and<br />

modulate transcriptional pathways to balance surfactant<br />

synthesis with surfactant secretion, recycling,<br />

and degradation is completely unknown.<br />

Likewise, it is unclear to what extent regulatory<br />

networks involved in type II cell maturation are<br />

also involved in alveolar repair and reconstitution<br />

of surfactant homeostasis following lung injury.<br />

These critical knowledge gaps are reflected in the<br />

paucity of new therapies for lung immaturity and<br />

chronic lung diseases. <br />

Research in the laboratories of the authors is currently<br />

supported by grants from Spanish Ministry of Science<br />

(BIO2009-09694, CSD2007-0010), Community of Madrid<br />

(S0505/MAT/0283), and Universidad Complutense to J.<br />

Perez-Gil and The National Heart Lung and Blood Institute<br />

(HL-056285 and HL-086492) to T. E. Weaver.<br />

References<br />

1. Almlen A, Stichtenoth G, Linderholm B, Haegerstrand-Bjorkman<br />

M, Robertson B, Johansson J, Curstedt T. <strong>Surfactant</strong><br />

proteins B and C are both necessary for alveolar stability at<br />

end expiration in premature rabbits with respiratory distress<br />

syndrome. J Appl Physiol 104: 1101–1108, 2008.<br />

2. Alpy F, Tomasetto C. Give lipids a START: the StAR-related<br />

lipid transfer (START) domain in mammals. J Cell Sci 118:<br />

2791–2801, 2005.<br />

3. Amin RS, Wert SE, Baughman RP, Tomashefski JF Jr, Nogee<br />

LM, Brody AS, Hull WM, Whitsett JA. <strong>Surfactant</strong> protein<br />

deficiency in familial interstitial lung disease. J Pediatr 139:<br />

85–92, 2001.<br />

4. Andreeva AV, Kutuzov MA, Voyno-Yasenetskaya TA. Regulation<br />

of surfactant secretion in alveolar type II cells. Am J<br />

Physiol Lung Cell Mol Physiol 293: L259–L271, 2007.<br />

5. Artemenko IP, Zhao D, Hales DB, Hales KH, Jefcoate CR.<br />

Mitochondrial processing of newly synthesized steroidogenic<br />

acute regulatory protein (StAR), but not total StAR, mediates<br />

cholesterol transfer to cytochrome P450 side chain cleavage<br />

enzyme in adrenal cells. J Biol Chem 276: 46583–46596,<br />

2001.<br />

6. Bachofen H, Gerber U, Gehr P, Amrein M, Schurch S. Structures<br />

of pulmonary surfactant films adsorbed to an air-liquid<br />

interface in vitro. Biochim Biophys Acta 1720: 59–72, 2005.<br />

7. Ban N, Matsumura Y, Sakai H, Takanezawa Y, Sasaki M, Arai<br />

H, Inagaki N. ABCA3 as a lipid transporter in pulmonary<br />

surfactant biogenesis. J Biol Chem 282: 9628–9634, 2007.<br />

8. Batenburg JJ. <strong>Surfactant</strong> phospholipids: synthesis and storage.<br />

Am J Physiol Lung Cell Mol Physiol 262: L367–L385,<br />

1992.<br />

9. Batenburg JJ, Haagsman HP. The lipids of pulmonary surfactant:<br />

dynamics and interactions with proteins. Prog Lipid Res<br />

37: 235–276, 1998.<br />

10. Batenburg JJ, Ossendorp BC, Snoek GT, Wirtz KW, Houweling<br />

M, Elfring RH. Phospholipid-transfer proteins and their<br />

mRNAs in developing rat lung and in alveolar type-II cells.<br />

Biochem J 298: 223–229, 1994.<br />

10a. Bernardino de la Serna J, Perez-Gil J, Simonsen AC, Bagatolli<br />

LA. Cholesterol rules: direct observation of the coexistence<br />

of two fluid phases in native pulmonary surfactant membranes<br />

at physiological temperatures. J Biol Chem 279:<br />

40715–40722, 2004.<br />

11. Bi X, Flach CR, Perez-Gil J, Plasencia I, Andreu D, Oliveira E,<br />

Mendelsohn R. Secondary structure and lipid interactions of<br />

the N-terminal segment of pulmonary surfactant SP-C in<br />

Langmuir films: IR reflection-absorption spectroscopy and<br />

surface pressure studies. Biochemistry 41: 8385–8395, 2002.<br />

12. Brasch F, Ochs M, Kahne T, Guttentag S, Schauer-Vukasinovic<br />

V, Derrick M, Johnen G, Kapp N, Muller KM, Richter J,<br />

Giller T, Hawgood S, Buhling F. Involvement of napsin A in<br />

the C- and N-terminal processing of surfactant protein B in<br />

type-II pneumocytes of the human lung. J Biol Chem 278:<br />

49006–49014, 2003.<br />

13. Brasch F, Schimanski S, Muhlfeld C, Barlage S, Langmann T,<br />

Aslanidis C, Boettcher A, Dada A, Schroten H, Mildenberger<br />

E, Prueter E, Ballmann M, Ochs M, Johnen G, Griese M,<br />

Schmitz G. Alteration of the pulmonary surfactant system in<br />

full-term infants with hereditary ABCA3 deficiency. Am J<br />

Respir Crit Care Med 174: 571–580, 2006.<br />

14. Bruder E, Hofmeister J, Aslanidis C, Hammer J, Bubendorf L,<br />

Schmitz G, Rufle A, Buhrer C. Ultrastructural and molecular<br />

analysis in fatal neonatal interstitial pneumonia caused by a<br />

novel ABCA3 mutation. Mod Pathol 20: 1009–1018, 2007.<br />

15. Bullard JE, Nogee LM. Heterozygosity for ABCA3 mutations<br />

modifies the severity of lung disease associated with a surfactant<br />

protein C gene (SFTPC) mutation. Pediatr Res 62:<br />

176–179, 2007.<br />

16. Butler PL, Mallampalli RK. Cross-talk between remodeling<br />

and de novo pathways maintains phospholipid balance<br />

through ubiquitination. J Biol Chem 285: 6246–6258, 2010.<br />

17. Cabre EJ, Malmstrom J, Sutherland D, Perez-Gil J, Otzen DE.<br />

<strong>Surfactant</strong> protein SP-B strongly modifies surface collapse of<br />

phospholipid vesicles: insights from a quartz crystal microbalance<br />

with dissipation. Biophys J 97: 768–776, 2009.<br />

18. Carnielli VP, Zimmermann LJ, Hamvas A, Cogo PE. Pulmonary<br />

surfactant kinetics of the newborn infant: novel insights<br />

from studies with stable isotopes. J Perinatol 29, Suppl 2:<br />

S29–S37, 2009.<br />

19. Clark JC, Wert SE, Bachurski CJ, Stahlman MT, Stripp BR,<br />

Weaver TE, Whitsett JA. Targeted disruption of the surfactant<br />

protein B gene disrupts surfactant homeostasis, causing<br />

respiratory failure in newborn mice. Proc Natl Acad Sci USA<br />

92: 7794–7798, 1995.<br />

19a. Clements JA. Lung surfactant: a personal perspective. Annu<br />

Rev Physiol 59: 1–21, 1997.<br />

20. Conkright JJ, Apsley KS, Martin EP, Ridsdale R, Rice WR, Na<br />

CL, Yang B, Weaver TE. NEDD4-2 mediated ubiquitination<br />

facilitates processing of surfactant protein C. Am J Respir<br />

Cell Mol Biol 2009.<br />

Downloaded from physiologyonline.physiology.org on September 9, 2010<br />

PHYSIOLOGY • Volume 25 • June 2010 • www.physiologyonline.org 139


REVIEWS<br />

21. Conkright JJ, Bridges JP, Na CL, Voorhout WF,<br />

Trapnell B, Glasser SW, Weaver TE. Secretion of<br />

surfactant protein C, an integral membrane protein,<br />

requires the N-terminal propeptide. J Biol<br />

Chem 276: 14658–14664, 2001.<br />

22. Cruz A, Casals C, Keough KM, Perez-Gil J. Different<br />

modes of interaction of pulmonary surfactant<br />

protein SP-B in phosphatidylcholine bilayers.<br />

Biochem J 327: 133–138, 1997.<br />

22a. Cruz A, Vazquez L, Velez M, Perez-Gil J. Effect of<br />

pulmonary surfactant protein SP-B on the microand<br />

nanostructure of phospholipid films. Biophys<br />

J 86: 308–320, 2004.<br />

23. Cruz A, Worthman LA, Serrano AG, Casals C,<br />

Keough KM, Perez-Gil J. Microstructure and dynamic<br />

surface properties of surfactant protein<br />

SP-B/dipalmitoylphosphatidylcholine interfacial<br />

films spread from lipid-protein bilayers. Eur Biophys<br />

J 29: 204–213, 2000.<br />

24. Chang R, Nir S, Poulain FR. Analysis of binding<br />

and membrane destabilization of phospholipid<br />

membranes by surfactant apoprotein B. Biochim<br />

Biophys Acta 1371: 254–264, 1998.<br />

25. Chen JW, Dodia C, Feinstein SI, Jain MK, Fisher<br />

AB. 1-Cys peroxiredoxin, a bifunctional enzyme<br />

with glutathione peroxidase and phospholipase<br />

A2 activities. J Biol Chem 275: 28421–28427,<br />

2000.<br />

26. Cheong N, Zhang H, Madesh M, Zhao M, Yu K,<br />

Dodia C, Fisher AB, Savani RC, Shuman H.<br />

ABCA3 is critical for lamellar body biogenesis in<br />

vivo. J Biol Chem 282: 23811–23817, 2007.<br />

27. Chevalier G, Collet AJ. In vivo incorporation of<br />

choline- 3 H, leucine- 3 H and galactose- 3Hin<br />

alveolar type II pneumocytes in relation to surfactant<br />

synthesis. A quantitative radoautographic<br />

study in mouse by electron microscopy.<br />

Anat Rec 174: 289–310, 1972.<br />

27a. de la Serna JB, Oradd G, Bagatolli LA, Simonsen<br />

AC, Marsh D, Lindblom G, Perez-Gil J. Segregated<br />

phases in pulmonary surfactant membranes<br />

do not show coexistence of lipid<br />

populations with differentiated dynamic properties.<br />

Biophys J 97: 1381–1389, 2009.<br />

28. Diemel RV, Snel MM, Waring AJ, Walther FJ, van<br />

Golde LM, Putz G, Haagsman HP, Batenburg JJ.<br />

Multilayer formation upon compression of surfactant<br />

monolayers depends on protein concentration<br />

as well as lipid composition. An atomic force microscopy<br />

study. J Biol Chem 277: 21179–21188,<br />

2002.<br />

29. Dietl P, Haller T. Exocytosis of lung surfactant:<br />

from the secretory vesicle to the air-liquid interface.<br />

Annu Rev Physiol 67: 595–621, 2005.<br />

30. Dietl P, Liss B, Felder E, Miklavc P, Wirtz H.<br />

Lamellar body exocytosis by cell stretch or purinergic<br />

stimulation: possible physiological roles,<br />

messengers and mechanisms. Cell Physiol Biochem<br />

25: 1–12, 2010.<br />

30a. Discher BM, Maloney KM, Schief WR, Jr.,<br />

Grainger DW, Vogel V, Hall SB. Lateral phase<br />

separation in interfacial films of pulmonary surfactant.<br />

Biophys J 71: 2583–2590, 1996.<br />

31. Eckenhoff RG, Somlyo AP. Rat lung type II cell<br />

and lamellar body: elemental composition in situ.<br />

Am J Physiol Cell Physiol 254: C614–C620, 1988.<br />

32. Edwards V, Cutz E, Viero S, Moore AM, Nogee L.<br />

Ultrastructure of lamellar bodies in congenital<br />

surfactant deficiency. Ultrastruct Pathol 29: 503–<br />

509, 2005.<br />

33. Filgueiras OM, Possmayer F. Purification and<br />

characterization of a phospholipase A2 associated<br />

with rabbit lung microsomes: some evidence<br />

for its mitochondrial origin. Biochim<br />

Biophys Acta 1046: 258–266, 1990.<br />

34. Fisher AB, Dodia C. Lysosomal-type PLA2 and<br />

turnover of alveolar DPPC. Am J Physiol Lung<br />

Cell Mol Physiol 280: L748–L754, 2001.<br />

35. Fitzgerald ML, Xavier R, Haley KJ, Welti R, Goss<br />

JL, Brown CE, Zhuang DZ, Bell SA, Lu N, McKee<br />

M, Seed B, Freeman MW. ABCA3 inactivation in<br />

mice causes respiratory failure, loss of pulmonary<br />

surfactant, and depletion of lung phosphatidylglycerol.<br />

J Lipid Res 48: 621–632, 2007.<br />

36. Follows D, Tiberg F, Thomas RK, Larsson M. Multilayers<br />

at the surface of solutions of exogenous<br />

lung surfactant: direct observation by neutron<br />

reflection. Biochim Biophys Acta 1768: 228–235,<br />

2007.<br />

37. Foster C, Aktar A, Kopf D, Zhang P, Guttentag S.<br />

Pepsinogen C: a type 2 cell-specific protease. Am<br />

J Physiol Lung Cell Mol Physiol 286: L382–L387,<br />

2004.<br />

38. Funkhouser JD, Read RJ. Phospholipid transfer<br />

proteins from lung, properties and possible physiological<br />

functions. Chem Phys Lipids 38: 17–27,<br />

1985.<br />

39. Garmany TH, Moxley MA, White FV, Dean M,<br />

Hull WM, Whitsett JA, Nogee LM, Hamvas A.<br />

<strong>Surfactant</strong> composition and function in patients<br />

with ABCA3 mutations. Pediatr Res 59: 801–805,<br />

2006.<br />

40. Glasser SW, Burhans MS, Korfhagen TR, Na CL,<br />

Sly PD, Ross GF, Ikegami M, Whitsett JA. Altered<br />

stability of pulmonary surfactant in SP-C-deficient<br />

mice. Proc Natl Acad Sci USA 98: 6366–<br />

6371, 2001.<br />

41. Glasser SW, Detmer EA, Ikegami M, Na CL,<br />

Stahlman MT, Whitsett JA. Pneumonitis and emphysema<br />

in sp-C gene targeted mice. J Biol<br />

Chem 278: 14291–14298, 2003.<br />

42. Gunasekara L, Schoel WM, Schurch S, Amrein<br />

MW. A comparative study of mechanisms of surfactant<br />

inhibition. Biochim Biophys Acta 1778:<br />

433–444, 2008.<br />

43. Gunther A, Schmidt R, Feustel A, Meier U, Pucker<br />

C, Ermert M, Seeger W. <strong>Surfactant</strong> subtype conversion<br />

is related to loss of surfactant apoprotein<br />

B and surface activity in large surfactant aggregates.<br />

Experimental and clinical studies. Am J<br />

Respir Crit Care Med 159: 244–251, 1999.<br />

44. Gustafsson M, Griffiths WJ, Furusjo E, Johansson<br />

J. The palmitoyl groups of lung surfactant protein<br />

C reduce unfolding into a fibrillogenic intermediate.<br />

J Mol Biol 310: 937–950, 2001.<br />

45. Guttentag S, Robinson L, Zhang P, Brasch F,<br />

Buhling F, Beers M. Cysteine protease activity is<br />

required for surfactant protein B processing and<br />

lamellar body genesis. Am J Respir Cell Mol Biol<br />

28: 69–79, 2003.<br />

46. Haagsman HP, Hogenkamp A, van Eijk M,<br />

Veldhuizen EJ. <strong>Surfactant</strong> collectins and innate<br />

immunity. Neonatology 93: 288–294, 2008.<br />

47. Haller T, Dietl P, Stockner H, Frick M, Mair N,<br />

Tinhofer I, Ritsch A, Enhorning G, Putz G. Tracing<br />

surfactant transformation from cellular release to<br />

insertion into an air-liquid interface. Am J Physiol<br />

Lung Cell Mol Physiol 286: L1009–L1015, 2004.<br />

48. Hammel M, Michel G, Hoefer C, Klaften M, Muller-Hocker<br />

J, de Angelis MH, Holzinger A. Targeted<br />

inactivation of the murine Abca3 gene<br />

leads to respiratory failure in newborns with defective<br />

lamellar bodies. Biochem Biophys Res<br />

Commun 359: 947–951, 2007.<br />

49. Hawgood S, Derrick M, Poulain F. Structure and<br />

properties of surfactant protein B. Biochim Biophys<br />

Acta 1408: 150–160, 1998.<br />

50. Ikegami M, Grant S, Korfhagen T, Scheule RK,<br />

Whitsett JA. <strong>Surfactant</strong> protein-D regulates the<br />

postnatal maturation of pulmonary surfactant<br />

lipid pool sizes. J Appl Physiol 106: 1545–1552,<br />

2009.<br />

51. Ikegami M, Na CL, Korfhagen TR, Whitsett JA.<br />

<strong>Surfactant</strong> protein D influences surfactant ultrastructure<br />

and uptake by alveolar type II cells. Am<br />

J Physiol Lung Cell Mol Physiol 288: L552–L561,<br />

2005.<br />

52. Johansson H, Eriksson M, Nordling K, Presto J,<br />

Johansson J. The Brichos domain of prosurfactant<br />

protein C can hold and fold a transmembrane<br />

segment. Protein Sci 18: 1175–1182,<br />

2009.<br />

53. Johansson H, Nordling K, Weaver TE, Johansson<br />

J. The Brichos domain-containing C-terminal part<br />

of pro-surfactant protein C binds to an unfolded<br />

poly-val transmembrane segment. J Biol Chem<br />

281: 21032–21039, 2006.<br />

54. Johnson AL, Braidotti P, Pietra GG, Russo SJ,<br />

Kabore A, Wang WJ, Beers MF. Post-translational<br />

processing of surfactant protein-C proprotein:<br />

targeting motifs in the NH 2 -terminal<br />

flanking domain are cleaved in late compartments.<br />

Am J Respir Cell Mol Biol 24: 253–263,<br />

2001.<br />

55. Kanno K, Wu MK, Scapa EF, Roderick SL, Cohen<br />

DE. Structure and function of phosphatidylcholine<br />

transfer protein (PC-TP)/StarD2. Biochim Biophys<br />

Acta 1771: 654–662, 2007.<br />

56. Keller A, Eistetter HR, Voss T, Schafer KP. The<br />

pulmonary surfactant protein C (SP-C) precursor<br />

is a type II transmembrane protein. Biochem J<br />

277: 493–499, 1991.<br />

57. Kingma PS, Whitsett JA. In defense of the lung:<br />

surfactant protein A and surfactant protein D.<br />

Curr Opin Pharmacol 6: 277–283, 2006.<br />

58. Kotorashvili A, Russo SJ, Mulugeta S, Guttentag<br />

S, Beers MF. Anterograde transport of surfactant<br />

protein C proprotein to distal processing compartments<br />

requires PPDY-mediated association<br />

with Nedd4 ubiquitin ligases. J Biol Chem 284:<br />

16667–16678, 2009.<br />

59. Kramer A, Wintergalen A, Sieber M, Galla HJ,<br />

Amrein M, Guckenberger R. Distribution of the<br />

surfactant-associated protein C within a lung<br />

surfactant model film investigated by near-field<br />

optical microscopy. Biophys J 78: 458–465,<br />

2000.<br />

60. Kuroki Y, Takahashi M, Nishitani C. Pulmonary<br />

collectins in innate immunity of the lung. Cell<br />

Microbiol 9: 1871–1879, 2007.<br />

61. Kuronuma K, Mitsuzawa H, Takeda K, Nishitani<br />

C, Chan ED, Kuroki Y, Nakamura M, Voelker DR.<br />

Anionic pulmonary surfactant phospholipids inhibit<br />

inflammatory responses from alveolar macrophages<br />

and U937 cells by binding the<br />

lipopolysaccharide-interacting proteins CD14<br />

and MD-2. J Biol Chem 284: 25488–25500, 2009.<br />

62. Lin S, Na CL, Akinbi HT, Apsley KS, Whitsett JA,<br />

Weaver TE. <strong>Surfactant</strong> protein B (SP-B) / mice<br />

are rescued by restoration of SP-B expression in<br />

alveolar type II cells but not Clara cells. J Biol<br />

Chem 274: 19168–19174, 1999.<br />

63. Lumb RH. Phospholipid transfer proteins in mammalian<br />

lung. Am J Physiol Lung Cell Mol Physiol<br />

257: L190–L194, 1989.<br />

64. Mao G, Desai J, Flach CR, Mendelsohn R. Structural<br />

characterization of the monolayer-multilayer<br />

transition in a pulmonary surfactant model:<br />

IR studies of films transferred at continuously<br />

varying surface pressures. Langmuir 24: 2025–<br />

2034, 2008.<br />

65. Mason RJ, Pan T, Edeen KE, Nielsen LD, Zhang F,<br />

Longphre M, Eckart MR, Neben S. Keratinocyte<br />

growth factor and the transcription factors<br />

C/EBP alpha, C/EBP delta, and SREBP-1c regulate<br />

fatty acid synthesis in alveolar type II cells. J<br />

Clin Invest 112: 244–255, 2003.<br />

66. Matsumura Y, Ban N, Ueda K, Inagaki N. Characterization<br />

and classification of ATP-binding cassette<br />

transporter ABCA3 mutants in fatal surfactant deficiency.<br />

J Biol Chem 281: 34503–34514, 2006.<br />

67. Melton KR, Nesslein LL, Ikegami M, Tichelaar JW,<br />

Clark JC, Whitsett JA, Weaver TE. SP-B deficiency<br />

causes respiratory failure in adult mice.<br />

Am J Physiol Lung Cell Mol Physiol 285: L543–<br />

L549, 2003.<br />

Downloaded from physiologyonline.physiology.org on September 9, 2010<br />

140<br />

PHYSIOLOGY • Volume 25 • June 2010 • www.physiologyonline.org


REVIEWS<br />

68. Mulugeta S, Gray JM, Notarfrancesco KL, Gonzales<br />

LW, Koval M, Feinstein SI, Ballard PL, Fisher<br />

AB, Shuman H. Identification of LBM180, a lamellar<br />

body limiting membrane protein of alveolar<br />

type II cells, as the ABC transporter protein<br />

ABCA3. J Biol Chem 277: 22147–22155, 2002.<br />

68a. Nag K, Perez-Gil J, Ruano ML, Worthman LA,<br />

Stewart J, Casals C, Keough KM. Phase transitions<br />

in films of lung surfactant at the air-water<br />

interface. Biophys J 74: 2983–2995, 1998.<br />

69. Nakanishi H, Shindou H, Hishikawa D, Harayama<br />

T, Ogasawara R, Suwabe A, Taguchi R, Shimizu T.<br />

Cloning and characterization of mouse lung-type<br />

acyl-CoA:lysophosphatidylcholine acyltransferase 1<br />

(LPCAT1). Expression in alveolar type II cells and<br />

possible involvement in surfactant production. J<br />

Biol Chem 281: 20140–20147, 2006.<br />

70. Nogee LM, de Mello DE, Dehner LP, Colten HR.<br />

Brief report: deficiency of pulmonary surfactant<br />

protein B in congenital alveolar proteinosis. N<br />

Engl J Med 328: 406–410, 1993.<br />

71. Nogee LM, Garnier G, Dietz HC, Singer L, Murphy<br />

AM, deMello DE, Colten HR. A mutation in<br />

the surfactant protein B gene responsible for<br />

fatal neonatal respiratory disease in multiple kindreds.<br />

J Clin Invest 93: 1860–1863, 1994.<br />

72. Numata M, Chu HW, Dakhama A, Voelker DR.<br />

Pulmonary surfactant phosphatidylglycerol inhibits<br />

respiratory syncytial virus-induced inflammation<br />

and infection. Proc Natl Acad Sci USA 107:<br />

320–325, 2010.<br />

73. Oosterlaken-Dijksterhuis MA, Haagsman HP, van<br />

Golde LM, Demel RA. Characterization of lipid<br />

insertion into monomolecular layers mediated by<br />

lung surfactant proteins SP-B and SP-C. Biochemistry<br />

30: 10965–10971, 1991.<br />

74. Oosterlaken-Dijksterhuis MA, Haagsman HP, van<br />

Golde LM, Demel RA. Interaction of lipid vesicles<br />

with monomolecular layers containing lung surfactant<br />

proteins SP-B or SP-C. Biochemistry 30:<br />

8276–8281, 1991.<br />

75. Osanai K, Mason RJ, Voelker DR. Pulmonary surfactant<br />

phosphatidylcholine transport bypasses<br />

the brefeldin A sensitive compartment of alveolar<br />

type II cells. Biochim Biophys Acta 1531: 222–<br />

229, 2001.<br />

76. Pastrana-Rios B, Flach CR, Brauner JW, Mautone<br />

AJ, Mendelsohn R. A direct test of the “squeezeout”<br />

hypothesis of lung surfactant function. External<br />

reflection FT-IR at the air/water interface.<br />

Biochemistry 33: 5121–5127, 1994.<br />

77. Patthy L. Homology of the precursor of pulmonary<br />

surfactant-associated protein SP-B with prosaposin<br />

and sulfated glycoprotein 1. J Biol Chem<br />

266: 6035–6037, 1991.<br />

78. Perez-Gil J. Structure of pulmonary surfactant<br />

membranes and films: the role of proteins and<br />

lipid-protein interactions. Biochim Biophys Acta<br />

1778: 1676–1695, 2008.<br />

79. Perez-Gil J, Keough KM. Interfacial properties of<br />

surfactant proteins. Biochim Biophys Acta 1408:<br />

203–217, 1998.<br />

80. Perez-Gil J, Nag K, Taneva S, Keough KM. Pulmonary<br />

surfactant protein SP-C causes packing<br />

rearrangements of dipalmitoylphosphatidylcholine<br />

in spread monolayers. Biophys J 63: 197–<br />

204, 1992.<br />

81. Perez-Gil J, Tucker J, Simatos G, Keough KM.<br />

Interfacial adsorption of simple lipid mixtures combined<br />

with hydrophobic surfactant protein from<br />

pig lung. Biochem Cell Biol 70: 332–338, 1992.<br />

83. Rodriguez-Capote K, Manzanares D, Haines T,<br />

Possmayer F. Reactive oxygen species inactivation<br />

of surfactant involves structural and functional<br />

alterations to surfactant proteins SP-B and<br />

SP-C. Biophys J 90: 2808–2821, 2006.<br />

84. Rooney SA, Young SL, Mendelson CR. Molecular<br />

and cellular processing of lung surfactant. FASEB<br />

J 8: 957–967, 1994.<br />

85. Ruano ML, Perez-Gil J, Casals C. Effect of acidic<br />

pH on the structure and lipid binding properties<br />

of porcine surfactant protein A. Potential role of<br />

acidification along its exocytic pathway. J Biol<br />

Chem 273: 15183–15191, 1998.<br />

86. Ryan MA, Qi X, Serrano AG, Ikegami M, Perez-Gil<br />

J, Johansson J, Weaver TE. Mapping and analysis<br />

of the lytic and fusogenic domains of surfactant<br />

protein B. Biochemistry 44: 861–872,<br />

2005.<br />

87. Saugstad OD, Hansen TW, Ronnestad A, Nakstad<br />

B, Tollofsrud PA, Reinholt F, Hamvas A,<br />

Coles FS, Dean M, Wert SE, Whitsett JA, Nogee<br />

LM. Novel mutations in the gene encoding ATP<br />

binding cassette protein member A3 (ABCA3)<br />

resulting in fatal neonatal lung disease. Acta Paediatr<br />

96: 185–190, 2007.<br />

88. Schurch S, Qanbar R, Bachofen H, Possmayer F.<br />

The surface-associated surfactant reservoir in the<br />

alveolar lining. Biol Neonate 67, Suppl 1: 61–76,<br />

1995.<br />

89. Serrano AG, Perez-Gil J. Protein-lipid interactions<br />

and surface activity in the pulmonary surfactant<br />

system. Chem Phys Lipids 141: 105–118,<br />

2006.<br />

90. Serrano AG, Ryan M, Weaver TE, Perez-Gil J.<br />

Critical structure-function determinants within<br />

the N-terminal region of pulmonary surfactant<br />

protein SP-B. Biophys J 90: 238–249, 2006.<br />

91. Shulenin S, Nogee LM, Annilo T, Wert SE, Whitsett<br />

JA, Dean M. ABCA3 gene mutations in newborns<br />

with fatal surfactant deficiency. N Engl J<br />

Med 350: 1296–1303, 2004.<br />

92. Taneva S, Keough KM. Pulmonary surfactant proteins<br />

SP-B and SP-C in spread monolayers at the<br />

air-water interface: II. Monolayers of pulmonary<br />

surfactant protein SP-C and phospholipids. Biophys<br />

J 66: 1149–1157, 1994.<br />

93. ten Brinke A, Vaandrager AB, Haagsman HP, Ridder<br />

AN, van Golde LM, Batenburg JJ. Structural<br />

requirements for palmitoylation of surfactant<br />

protein C precursor. Biochem J 361: 663–671,<br />

2002.<br />

94. Tian Y, Zhou R, Rehg JE, Jackowski S. Role of<br />

phosphocholine cytidylyltransferase alpha in lung<br />

development. Mol Cell Biol 27: 975–982, 2007.<br />

95. Ueda T, Ikegami M, Jobe <strong>Surfactant</strong> subtypes A.<br />

In vitro conversion, in vivo function, and effects<br />

of serum proteins. Am J Respir Crit Care Med<br />

149: 1254–1259, 1994.<br />

96. Ueno T, Linder S, Na CL, Rice WR, Johansson J,<br />

Weaver TE. Processing of pulmonary surfactant<br />

protein B by napsin and cathepsin H. J Biol Chem<br />

279: 16178–16184, 2004.<br />

97. van Helvoort A, de Brouwer A, Ottenhoff R,<br />

Brouwers JF, Wijnholds J, Beijnen JH, Rijneveld<br />

A, van der Poll T, van der Valk MA, Majoor D,<br />

Voorhout W, Wirtz KW, Elferink RP, Borst P. Mice<br />

without phosphatidylcholine transfer protein<br />

have no defects in the secretion of phosphatidylcholine<br />

into bile or into lung airspaces. Proc Natl<br />

Acad Sci USA 96: 11501–11506, 1999.<br />

100. Veldhuizen R, Nag K, Orgeig S, Possmayer F. The<br />

role of lipids in pulmonary surfactant. Biochim<br />

Biophys Acta 1408: 90–108, 1998.<br />

101. von Nahmen A, Post A, Galla HJ, Sieber M. The<br />

phase behavior of lipid monolayers containing<br />

pulmonary surfactant protein C studied by fluorescence<br />

light microscopy. Eur Biophys J 26:<br />

359–369, 1997.<br />

102. von Nahmen A, Schenk M, Sieber M, Amrein M.<br />

The structure of a model pulmonary surfactant as<br />

revealed by scanning force microscopy. Biophys<br />

J 72: 463–469, 1997.<br />

103. Wade KC, Guttentag SH, Gonzales LW, Maschhoff<br />

KL, Gonzales J, Kolla V, Singhal S, Ballard PL.<br />

Gene induction during differentiation of human<br />

pulmonary type II cells in vitro. Am J Respir Cell<br />

Mol Biol 34: 727–737, 2006.<br />

104. Wang L, Magdaleno S, Tabas I, Jackowski S. Early<br />

embryonic lethality in mice with targeted deletion<br />

of the CTP:phosphocholine cytidylyltransferase<br />

alpha gene (Pcyt1a). Mol Cell Biol 25:<br />

3357–3363, 2005.<br />

105. Wang P, Chintagari NR, Narayanaperumal J, Ayalew<br />

S, Hartson S, Liu L. Proteomic analysis of<br />

lamellar bodies isolated from rat lungs. BMC Cell<br />

Biol 9: 34, 2008.<br />

106. Wang Z, Gurel O, Baatz JE, Notter RH. Differential<br />

activity and lack of synergy of lung surfactant<br />

proteins SP-B and SP-C in interactions with<br />

phospholipids. J Lipid Res 37: 1749–1760,<br />

1996.<br />

107. Whitsett JA, Wert SE, Weaver TE. Alveolar surfactant<br />

homeostasis and the pathogenesis of pulmonary<br />

disease. Annu Rev Med 61: 105–119.<br />

108. Wirtz KW. Phospholipid transfer proteins in perspective.<br />

FEBS Lett 580: 5436–5441, 2006.<br />

109. Wright JR. Immunoregulatory functions of surfactant<br />

proteins. Nat Rev Immunol 5: 58–68, 2005.<br />

110. Wu G, Aoyama C, Young SG, Vance DE. Early<br />

embryonic lethality caused by disruption of the<br />

gene for choline kinase alpha, the first enzyme in<br />

phosphatidylcholine biosynthesis. J Biol Chem<br />

283: 1456–1462, 2008.<br />

111. Yamano G, Funahashi H, Kawanami O, Zhao LX,<br />

Ban N, Uchida Y, Morohoshi T, Ogawa J, Shioda<br />

S, Inagaki N. ABCA3 is a lamellar body membrane<br />

protein in human lung alveolar type II cells.<br />

FEBS Lett 508: 221–225, 2001.<br />

112. Yang L, Johansson J, Ridsdale R, Willander H,<br />

Fitzen M, Akinbi HT, Weaver TE. <strong>Surfactant</strong> protein<br />

B propeptide contains a saposin-like protein<br />

domain with antimicrobial activity at low pH. J<br />

Immunol 184: 975–983.<br />

113. Zimmermann LJ, Janssen DJ, Tibboel D, Hamvas<br />

A, Carnielli VP. <strong>Surfactant</strong> metabolism in the neonate.<br />

Biol Neonate 87: 296–307, 2005.<br />

114. Zuo YY, Veldhuizen RA, Neumann AW, Petersen<br />

NO, Possmayer F. Current perspectives in pulmonary<br />

surfactant: inhibition, enhancement and<br />

evaluation. Biochim Biophys Acta 1778: 1947–<br />

1977, 2008.<br />

115. Zuo YY, Keating E, Zhao L, Tadayyon SM,<br />

Veldhuizen RA, Petersen NO, Possmayer F.<br />

Atomic force microscopy studies of functional<br />

and dysfunctional pulmonary surfactant films. I.<br />

Micro- and nanostructures of functional pulmonary<br />

surfactant films and the effect of SP-A. Biophys<br />

J 94: 3549–3564, 2008.<br />

Downloaded from physiologyonline.physiology.org on September 9, 2010<br />

82. Postle AD, Heeley EL, Wilton DC. A comparison<br />

of the molecular species compositions of mammalian<br />

lung surfactant phospholipids. Comp Biochem<br />

Physiol A Mol Integr Physiol 129: 65–73, 2001.<br />

98. van Meer G, Voelker DR, Feigenson GW. Membrane<br />

lipids: where they are and how they behave.<br />

Nat Rev Mol Cell Biol 9: 112–124, 2008.<br />

99. Veldhuizen EJ, Batenburg JJ, van Golde LM,<br />

Haagsman HP. The role of surfactant proteins in<br />

DPPC enrichment of surface films. Biophys J 79:<br />

3164–3171, 2000.<br />

PHYSIOLOGY • Volume 25 • June 2010 • www.physiologyonline.org 141

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!