Literature
Home医源资料库在线期刊美国呼吸和危急护理医学2005年第171卷第6期

Cellular Stress Failure in Ventilator-injured Lungs

来源:美国呼吸和危急护理医学
摘要:CellularStressFailureinInjuredLungsTotestifplasmamembraneinjuryandrepairarephenomenainventilator-injuredlungs,Gajicandcoworkersperfusedexvivomechanicallyventilatedratlungswithsolutionscontainingthemembrane-impermeantlabelpropidiumiodide(PI)(95)。CellularRem......

点击显示 收起

    Thoracic Diseases Research Unit, Division of Pulmonary and Critical Care Medicine, Department of Medicine
    Department of Physiology and Biomedical Engineering, Mayo Clinic College of Medicine, Rochester, Minnesota

    ABSTRACT

    The clinical and experimental literature has unequivocally established that mechanical ventilation with large tidal volumes is injurious to the lung. However, uncertainty about the micromechanics of injured lungs and the numerous degrees of freedom in ventilator settings leave many unanswered questions about the biophysical determinants of lung injury. In this review we focus on experimental evidence for lung cells as injury targets and the relevance of these studies for human ventilator-associated lung injury. In vitro, the stress-induced mechanical interactions between matrix and adherent cells are important for cellular remodeling as a means for preventing compromise of cell structure and ultimately cell injury or death. In vivo, these same principles apply. Large tidal volume mechanical ventilation results in physical breaks in alveolar epithelial and endothelial plasma membrane integrity and subsequent triggering of proinflammatory signaling cascades resulting in the cytokine milieu and pathologic and physiologic findings of ventilator-associated lung injury. Importantly, though, alveolar cells possess cellular repair and remodeling mechanisms that in addition to protecting the stressed cell provide potential molecular targets for the prevention and treatment of ventilator-associated lung injury in the future.

    Key Words: alveolar epithelium  cell injury  cell mechanics  cell repair  mechanical ventilation  plasma membrane tension

    INTRODUCTION

    In the United States over 100,000 individuals each year develop Acute Lung Injury (ALI) and Acute Respiratory Distress Syndrome (ARDS), the more severe form of ALI (1). Current estimates of attributable mortality for ALI/ARDS range between 17,000 and 43,000 persons per year. Numerous studies suggest that ALI mortality has improved during the past two decades, but remains high, ranging between 25 and 50% in the most current series (2eC6). While mechanical ventilation is integral to the management of patients with injured lungs, a recent ARDS network trial has made it clear that the choice of ventilator settings accounts for as many as one-third of all deaths attributed to ALI (7). This, and a preceding smaller trial (8), represent the successful bench-to-bedside culmination of decades of research that had suggested artificial ventilation can either cause de novo lung injury or aggravate preexisting lung injury (reviewed in Reference 9).

    The breadth and depth of knowledge regarding mechanical ventilation, ventilator-induced lung injury (VILI), barotrauma, "so-called" biotrauma, and mechanotransduction is far too great to do it justice in a single review. The reader is referred to several outstanding reviews on the pathophysiology and clinical manifestations of VILI (9, 10), on the molecular biology of pulmonary mechanotransduction (11eC18), and on the effects of deforming stress on surfactant biology (19eC22). In this review we will therefore focus primarily on the structural failure of cells and tissues, a relatively novel and underappreciated area of lung biology, as it is causally related to many of the disease manifestations in ventilator-injured lungs. Widespread endothelial and epithelial cell disruption and plasma membrane blebbing are indeed hallmarks of the entity and account at least in part for the increased microvascular permeability that is readily observed in experimental models of VILI (23eC28).

    PULMONARY MICROMECHANICS

    A review of the cellular pathology of ventilator-injured lungs would be incomplete without some discussion of the current understanding of pulmonary micromechanics in health and disease. It is important to note that there remain major gaps in knowledge because the available imaging tools lack sufficient temporal and/or spatial resolution to quantify stresses and strains on the scale of interest. Nonetheless, an appreciation of the governing principles provides the foundation for the design of physiologically relevant in vitro and in vivo studies, the physiologic basis for understanding experimental results, and ultimately a basis for their logical application to modes of clinical practice. Throughout this review we will use terms such as stress, strain, elastic modulus, or stiffness, and have listed their definitions in Table 1.

    Alveolar Micromechanics of the Normal Lung

    For more than 50 years it has been appreciated that the topographical distributions of lung parenchymal stress and strain are nonuniform, and the biophysical determinants of this nonuniformity are generally understood (29eC32). Accordingly, the lungs and the boundary structures to which they must conform (ribcage, diaphragm abdomen, heart, and mediastinum) are considered gravitationally deformed elastic solids. The shape matching of lung and boundary structures imposes a nonuniform stress and strain field. At least in quadrupeds, the weight of the lung is only a minor determinant of nonuniform transpulmonary pressures and alveolar volumes (30). The gravitational deformations of heart and diaphragm/abdomen turn out to be much more important determinants of regional volume and ventilation than the actions of gravity on the lungs themselves (33, 34). Furthermore, with increasing precision of methods for measuring regional lung function it is now apparent that there is considerable small-scale heterogeneity in lung parenchymal strain that cannot be explained by any gravitational mechanism (35, 36).

    Measurements of lung strain in recumbent dogs suggested that the linear dimensions of the lung increase by as much as 40% during an inflation from functional residual capacity (FRC) to total lung capacity (TLC) (37). However, this value is a gross overestimate of the elastic deformation experienced by individual cells and tissue matrix. The lung parenchyma is a connective tissue network that is distorted by surface tension (38). Embedded in this network are airways and blood vessels, which resist deformation to a greater extent than the surrounding parenchyma. This difference in mechanical properties is an important source of interdependence and explains why in cases of barotrauma extraalveolar air generally tracks along bronchovascular bundles or why edema fluid accumulates in perivascular cuffs (39). Models of parenchymal micromechanics that are based on morphometric analyses of perfusion-fixed tissue specimens consider the helical network of elastin and collagen fibers that form the alveolar ducts as the primary stress-bearing structures (40eC45). The alveolar walls in turn are largely supported by surface tension, and are thought to simply unfold as lung volume increases (46). This explains why macroscopic strains computed from lung regions that are more than 1 cm3 may grossly overestimate the stretch experienced by lung cells during breathing. Aware of this limitation, Tschumperlin and Margulies traced the lengths of alveolar basement membranes in electron microscopic images of alveolar walls and estimated their area change with transpulmonary pressure and volume (47). Accordingly, the basement membrane area increased by approximately 35% during an inspiratory capacity maneuver, which corresponds to a linear strain and hence cell stretch estimate of about 15% (47). These observations were in keeping with earlier work suggesting that in the normal tidal breathing range alveolar septae simply unfold as opposed to being stretched (38). Because most fixatives affect tissue hydration and surface tension and thereby distort lung architecture relative to its in vivo state, the current models of alveolar micromechanics await morphometric confirmation on living, unfixed specimens (48). Microscopic imaging of canine subpleural alveoli through a pleural window suggested that alveolar volume changes little during normal breathing (49) and that the acinus expands nonuniformly (43). Because the mechanics of subpleural alveoli may be dominated by their coupling to a relatively inelastic pleural membrane, which in these experiments had to be immobilized to generate a focused image, the amplitude of alveolar volume change during quiet breathing is likely to remain a topic of active investigation.

    Alveolar Micromechanics in Injury States

    There is a relative paucity of detailed morphometric data on injured lungs, and their interpretation is controversial (50eC55). The long-held view that the heavy injured lung collapses under its own weight has been challenged (56eC58). The challenge rests on the assertion that fluid accumulates in small airways and distal airspaces, which prevents as opposed to promotes the collapse of dependent lung tissue. The effects on gas exchange, i.e., shunt and low /, are similar, regardless of whether one views the dependent lung as airless and collapsed (tissue dimensions are decreased) or occluded by liquid plugs and expanded by edema (tissue dimensions are normal or increased). However, the stresses to which airway and alveolar lining cells are exposed during breathing could be quite different.

    Injured lungs possess two attributes that explain why they are at increased risk for additional deformation injury. The first attribute is that the number of alveoli capable of expanding during inspiration is decreased. The injured lung contains normally aerated, poorly aerated, and nonaerated respiratory units (59). A smaller number of less injured units are preferentially recruited (referred to by Gattinoni as "baby lung" [60]), and thus receive a large proportion of the delivered tidal volume. This explains the increased risk of injury from regional overexpansion. The second attribute is that the local impedance to lung expansion is heterogeneous and a result of the distribution of liquid and surface tension in distal airspaces. This heterogeneity in lung impedances results in shear stress being generated between neighboring, interdependent units that operate at different volumes, as was first detailed by Mead and colleagues (61). They pointed out that forces carried by elements of a uniformly expanded network structure must be uniformly distributed, but that departures from uniformity do generate large stress concentrations. One of the reasons for such stress concentrations, they argued, was that a change in dimension of one element (e.g., a collapsed alveolus) mandates that insertion forces of neighboring elements (the surrounding parenchyma) act over a smaller area. Consequently, local stress, defined as force per unit area, has to increase. Even though Mead's numeric analyses focused on the reduction in alveolar surface area as principle source of heterogeneous stress and ignored strain gradients in the surrounding parenchyma, the resulting insights have proven fundamental to understanding the mechanics of injured lungs.

    In addition to stress concentrations resulting from interdependence, there is also injury to small airways and alveolar ducts caused by their repeated opening and closure and by energy dissipation during liquid bridge fracture or stress that is imposed on lining cells by the movement of aireCliquid interfaces with respiration (62eC64). Modeling approaches to bubble and liquid flow in tubes, while constrained by simplifying assumptions (e.g., rigid tube of uniform diameter, smooth surface), are beginning to shed some light on more quantitative aspects of this problem (65, 66). The relative contribution of these distinct injury mechanisms in different syndromes and disease models is simply not known. Inferences from animal experiments with short-term physiologic endpoints are at best hypothesis generating, but have yet to demonstrate the circumstance under which any one of these injury mechanisms prevails.

    THE BLOODeCGAS BARRIER OF VENTILATOR-INJURED LUNGS

    The first experimental study of VILI appeared in 1964 and demonstrated that mechanical ventilation with high volumes and pressures altered the surface properties of canine lung extracts (67). In 1974 Webb and Tierney reported that mechanical ventilation with large tidal volumes caused hemorrhagic pulmonary edema in rats (68). The findings established that deforming stresses associated with mechanical ventilation could alter lung barrier function and moreover impair the integrity of the bloodeCgas barrier. The clinical relevance of these findings was not appreciated until Egan, and later Parker and coworkers, began to explore the effects of lung volume and mechanical ventilation on pulmonary vascular barrier properties (69eC73). The critical care community took note of this work only after Dreyfuss and colleagues confirmed Webb and Tierney's observations and demonstrated that tidal volume was a more appropriate determinant of deforming and potentially injurious stress than was peak airway pressure (reviewed in References 9 and 24).

    Morphology

    Dreyfuss and colleagues were the first to characterize in detail the morphology of the bloodeCgas barrier of ventilator-injured rat lungs (24), extending earlier observations by John and colleagues in mechanically ventilated rabbit lungs (27). Electron micrographs of rat lungs taken after 5 minutes of injurious ventilation showed interstitial edema and endothelial lesions consisting of plasma-membrane blebs and loss of cell contact with the basement membrane. More prolonged exposure to injurious stress produced alveolar epithelial pathology ranging in spectrum from inter- and intracellular gap formations (Figure 1A) with denuded basement membranes to extensive cell destruction (9, 24, 27). These changes in cellular ultrastructure may be viewed as evidence for deformation related cell remodeling and/or yielding of the cells' stress-bearing elements. Interestingly, type II alveolar epithelial cells appeared relatively spared, suggesting that they had experienced a smaller deformation on account of their location in alveolar corners.

    In a series of studies motivated by interest in high-altitude physiology, the group of West (reviewed in Reference 74) studied the consequences of capillary pressure on the bloodeCgas barrier (Figure 1B). The bloodeCgas barrier of rabbit lungs exposed to high capillary pressures revealed not only transcellular epithelial gaps and endothelial lesions, but also basement membrane breaks (75, 76). However, frank cell necrosis and large alveolar wounds were not observed. On the basis of these findings West's group suggested that under certain conditions capillary pressures could exceed the structural limit of the basement membrane, which is the primary stress-bearing element of the bloodeCgas barrier. For the most part it is composed of a network of type IV collagen fibers, which can withstand considerable tensile stress (74). Physiologic and pathologic conditions in which pulmonary capillary stress failure has been observed or strongly suspected include high-altitude pulmonary edema (77), congestive heart failure, mitral stenosis, and Goodpasture's Syndrome, which is characterized by an immune mediated weakening of the collagen IV lattice (78), as well as high intensity exercise in race horses (75) and elite athletes (79). West's group also emphasized important mechanical interactions between lung volume, capillary pressure, and the probability of capillary stress failure, which is in keeping with experimental observations on isolated perfused and mechanically ventilated rabbit lungs and a case report of a patient with ARDS (75, 80).

    Although structural failure of capillary basement membranes has not to date been demonstrated in ventilator-injured lungs, the presence of pulmonary hemorrhage in rat and canine VILI models, which is readily apparent to the naked eye, would be hard to explain by any other mechanism (24, 68, 81). At the same time, not all hydrostatic pulmonary edema results in capillary stress failure (53, 54). That is because the transmural pressures at which intraalveolar capillaries experience yield stress is quite high and was estimated by West and colleagues to approximate 40 mm Hg (75). Capillary pressures associated with ultrastuctural changes in adherent endothelial and epithelial cells tend to be considerably lower (54).

    Majno and coworkers recognized as early as 1969 that vascular permeability was at least in part controlled by contractile endothelial cell proteins (82). Majno and colleagues postulated that inflammatory mediators caused active endothelial contraction with formation of intercellular gaps and subsequent extravasation of plasma. Since then, a great deal has been learned about the endothelial regulation of pulmonary vascular barrier properties and about the role of adhesion receptors and cytoskeletal proteins in this process (reviewed in References 83eC87). Moreover, a body of work including the pioneering studies by Neal and Michel on frog mesenteric vessels (Figure 1C) has established that certain endothelial agonists and high vascular pressures cause gaps not only between adjacent endothelial cells but also within or through individual endothelial cells (88eC94). These gaps close rapidly upon removal of the deforming stress, restoring normal vascular permeability. These observations are in keeping with the observed plasticity of the bloodeCgas barrier in transient pulmonary venous hypertension (93) and intermittent hyperinflation (94).

    Cellular Stress Failure in Injured Lungs

    To test if plasma membrane injury and repair are phenomena in ventilator-injured lungs, Gajic and coworkers perfused ex vivo mechanically ventilated rat lungs with solutions containing the membrane-impermeant label propidium iodide (PI) (95). When PI enters a cell through a plasma membrane defect, it interchelates with DNA and emits red fluorescence upon excitation with blue light. PI-positive cells can therefore be identified in optical sections of subpleural airspaces obtained with laser confocal microscopy (Figure 2). In a series of validation experiments, Gajic and colleagues showed that the number of subpleural cells with membrane defects increases with increasing tidal volumes and duration of stress exposure, and that cell injury correlates reasonably well with other physiologic and histologic injury markers. More importantly, by comparing preparations that had been labeled during ventilation at injurious settings with those labeled after removal of the injurious stress, Gajic and associates inferred that over 60% of injured cells repair plasma membrane defects.

    In aggregate experimental studies confirm that injurious mechanical ventilation produces stress failure of capillary basement membranes as well as of adherent cells. It is not clear if epithelial and endothelial cells are the only lung cells predisposed to stress failure, what role the subcortical cytoskeleton plays in strain-related plasma membrane breaks, or if stress failure of the basement membrane invariably leads to stress failure of cells. It is also not clear if intercellular and/or intracellular gap formation in adherent cells is an active remodeling response to adapt to large substratum strains. In frog mesenteric microvessel experiments relevant to lung deformation, the pressure at which plasma extravasates through intracellular gaps was shown to be temperature-sensitive (96). Because low temperature had no measurable effect on the vessels' compliance, the investigators attributed gap formation to inhibition of active deformation-induced cell remodeling as opposed to structural failure consequent to a basement membrane break (i.e., the capillary is more leaky but not more fragile). Alternatively, recent experiments using rat pulmonary microvascular endothelial cells suggest that vessel leakiness resulting from mechanical cell wounding may be a result of weakened celleCcell adhesion resulting from decreases in the expression of the cell juctional protein, -catenin (97).

    CELL RESPONSES AND CONSEQUENCES OF DEFORMING STRESS

    The response of cells to deforming forces is a result of the cell's ability to "sense" and transduce these stimuli. Experiments studying the molecular mechanisms of mechanotransduction have implicated numerous candidates. However, the nature of their interrelatedness, cooperativity, and cell and tissue specificity remain areas of continued rigorous investigation. The reader is referred to a number of excellent reviews addressing cellular mechanotransduction (98eC100). In this section we will discuss the general principles of cellular microrheology as a platform for highlighting the cellular consequences and adaptive and reparative responses to cell deformation that mimics the injurious effect of mechanical ventilation in vivo.

    Microrheology of Living Cells

    The principle stress-bearing elements of the lung, which account for its tendency to recoil, are elastin and collagen fiber networks and surface tension. Indeed, the lung can be viewed as a tissue network that is distorted by surface tension (46). While the resistance of cells to deformation contributes little to overall lung stiffness, lung cells must nevertheless adapt to deformations of the scaffolding to which they adhere. Cells interact with their surroundings through adhesion receptors such as integrins, which provide dynamic bidirectional links between the cytoskeleton and the extracellular matrix (101eC103). An increase in basement membrane surface area that accompanies a large tidal breath thus imposes a shape change on adherent alveolar epithelial and microvascular endothelial cells. Both epithelial and endothelial cells are subjected to deforming stress during breathing as a result of the interdependent effects of lung volume, transpulmonary pressure, surface tension, and vascular pressure on the bloodeCgas barrier (75). The resulting cellular shape change mandates that cell surface to volume ratio increase, and this is generally accompanied by a reorganization of the cell's stress-bearing elements (i.e., the cytoskeleton).

    The cytoskeleton is an interconnected network of biopolymers that exert centripetal forces on the surrounding matrix (104, 105). It is covered by the plasma membrane, a lipid bilayer, the molecular constituents of which are organized in specific outer and inner leaflet domains (106). Compared with cytoskeletal proteins, the plasma membrane carries little stress under physiologic conditions (107). Nevertheless, it may experience lytic tensions when a large shape change is externally imposed. This can be readily documented in epithelial monolayers that are grown on malleable membranes and subjected to large deformations (108eC112). Interestingly, plasma membrane defects resulting from large substratum strains tend to be transient and rapidly repaired (112).

    The mechanical properties of solid materials can be described by elastic constants that characterize the material's resistance to changes in volume and shape (31). Within this framework, yield-stress and lytic stress are quantities denoting the material's susceptibility to plastic deformation and structural failure. Material properties have been estimated for cell constituents such as cytoskeletal proteins, cytoskeletal networks, lipid bilayers, and plasma membranes, as well as whole cells including alveolar and bronchial epithelial cells (113, 114). In the context of a discussion on cell injury two characteristics of biomaterials deserve comment: (1) the distinct rheologic properties of network structures (115eC117) and (2) the importance of active remodeling in determining cell plasticity (118eC120).

    Because cells and specifically their cytoskeleton are network structures, the mechanical properties of individual stress-bearing elements (e.g., single actin fibers) are only secondary determinants of a cell's deformation resistance. Cells are prestressed networks of tension-bearing microfilaments that are coupled to compression-resistant microtubules and extracellular matrix molecules (117). As such, their resistance to deformation is critically dependent on the interconnectedness of the network structure and on the rate at which molecular contacts between stress-bearing elements can be broken, degraded, and reformed. Living cells display a great deal of plasticity, that is, the network of stress-bearing elements remodels readily when it is subjected to a deforming stress (121eC123) Deformation induced remodeling involves active, energy-dependent processes that may well provide a safeguard against the development of structural failure (112, 124, 125).

    Cellular Remodeling: Prevention of Plasma Membrane Wounding

    Matrix- and cytoskeleton-dependent mechano-sensing and remodeling.

    Proteins that link the extracellular matrix with stress bearing cytoskeletal biopolymers play a pivotal role in mechanosensing and transduction (101, 126eC129). They are assembled in so-called focal adhesion sites, plaques, or complexes and can be thought of as mini-strain gauges that monitor local force (Figure 3). Focal adhesions are highly plastic structures and remodel in a force-dependent manner. This allows adherent cells to probe the regional impedance of the surrounding scaffolding and in turn provides cues for directional control of locomotion and cell shape (101, 105). The most extensively studied adhesion receptors involved in mechanotransduction are the integrins. After ligand-induced activation, integrins transduce matrix-dependent intracellular biochemical signals (130) through structural associations with adaptor proteins, which include GTPases (131), receptor and nonreceptor tyrosine kinases (132eC136), and phosphoinositides (137, 138). Other adhesion molecules such as cadherins are also increasingly recognized as playing important roles in mechanotransduction (133) and as molecular targets of deformation induced impairments of epithelial and endothelial barrier function (97, 139, 140).

    Through numerous distinct but interrelated biochemical signals, matrix molecules undoubtedly play a major role in lung remodeling induced by stress (141eC143). For example, bronchial epithelial cells exposed to a compressive stress in vitro elicit a profibrotic response in unstressed fibroblasts (144). Compressive stress was shown to shrink the lateral intercellular space surrounding epithelial cells, and thereby triggered signaling via autocrine binding of epidermal growth factor family ligands to the epidermal growth factor receptor (145). Remodeling responses are also initiated through paracrine signaling involving fibronectin, collagen, and matrix metalloproteinases (MMP). In a rat model of VILI, expression of the extracellular matrix metalloproteinase inducer (EMMPRIN), gelatinase A and B, MT1-MP were induced in lung endothelium (146, 147). In the rabbit, mechanical ventilation with high positive end-expiratory pressure (PEEP) was associated with increased mRNA expression of extracellular matrix proteins such as 2(IV) procollagen and fibronectin (141). These effects were measured in the absence of alveolar cell breaks, as quantified by electron microscopy, and were accompanied by an increase in mRNA of the mitogenic growth factors TGF-1 and basic fibroblast growth factor. The authors attributed the findings to wall stresseCinduced vascular remodeling.

    It is not our intent to provide a comprehensive review or even a complete list of the large number of publications dealing with injured lungs and matrix and adhesion molecules. We simply underscore that these molecules are integral to cell and tissue remodeling regardless of whether their expression is triggered by the structural failure of a stress-bearing element or initiated by some other mechanotransduction event.

    Regulation of cell surface area and plasma membrane tension.

    The hypothesis that deformation induced remodeling is vital in the prevention of cellular stress failure applies not only to remodeling responses involving matrix and cytoskeletal proteins, but also to vesicular lipid trafficking to and from the plasma membrane (Figure 3). This lipid trafficking serves as a means of regulating cell surface area and plasma membrane tension and ultimately helps to prevent plasma membrane stress failure (112, 124, 125, 148). To the extent to which plasma membrane stress failure is one of the triggers of proinflammatory signaling and/or cell death in ventilator-injured lungs, it becomes important to understand how cells regulate plasma membrane surface area and tension.

    In a thought-provoking essay, Morris discusses the evolutionary question of how apparent regulatory feedback loops for cell volume and surface area came into being (149). She concludes that size regulation in the earliest protocells would have been governed by liposome physics and develops the argument that monitoring and regulation of lipid bilayer tension ultimately determines a cell-size set point. According to that theory, changes in bilayer tension could have altered membrane conductivity for osmolytes and consequently effected a cell volume change. Modern cells have evolved more elaborate control mechanisms linked to protein-regulated expenditures of energy and fluxes of materials. Cells sense a multitude of flux rates such as the rates of endocytosis, exocytosis, protein channel or pump activities, and the polymerization rates of cytoskeletal networks. Volume, surface area, and shape are simply consequences of the weighted distributions of the respective rate constants, but are not the sensed quantities themselves. Plasma membrane tension is central to many of these control loops. As detailed below, changes in plasma membrane tension not only effect membrane ion and water conductivities but also the rates of membrane addition (exocytosis) and retrieval (endocytosis) (150). For example, increases in plasma membrane tension, in response to a hypoosmolar challenge, produce a net increase in plasma membrane, whereas a fall in tension results in membrane retrieval (151, 152).

    Deformation-induced lipid trafficking.

    Changes in the dimension of the connective tissue matrix during breathing impose a shape change in adherent cells which ought to intermittently raise plasma membrane tension. Therefore, Vlahakis and colleagues reasoned that to prevent lytic membrane tensions, alveolar epithelial cells would demonstrate a net exocytic lipid trafficking response during stretch in culture (125). Not only did they demonstrate active lipid vesicle trafficking to the plasma membrane that was temperature- and energy-dependent, but also that the response could be inhibited by both cytoskeleton active agents and by cholesterol depletion of the plasma membrane (112). Recent observations have confirmed the importance of deformation-induce lipid trafficking over plasma membrane unfolding in the surface area regulation of tonically stretched alveolar epithelial cells (148). It must be emphasized that the mechanisms underlying this trafficking response cannot be equated with those governing deformation-induced surfactant secretion by type II alveolar epithelial cells (153, 154).

    Deforming stresses can also trigger endocytic responses, as was demonstrated in umbrella cells of pressurized porcine bladders and in radially strained alveolar epithelial cells (155, 156). That stretch would trigger endocytosis of surface membrane lipids might seem counterintuitive when considered in the context of plasma membrane tension regulation. However, the experimental evidence is undeniable and suggests that exocytic and endocytic trafficking responses are intrinsically linked (157). While exocytosis and endocytosis may be temporarily dissociated, their relative rates probably do vary with plasma membrane tension. Furthermore, preliminary evidence from alveolar epithelial cell lines suggests that stretch-induced internalization of lipids proceeds via distinct molecular pathways (156).

    Several observations on normal and injured lungs raise interest in the molecule and pathway specificity of deformation triggered vesicular trafficking. When lungs suffer relatively mild forms of interstitial pulmonary edema, the lipid microdomains of lung cell surface membranes undergo a substantial reorganization (158, 159). The functional consequence of membrane remodeling, which is almost certainly accompanied by changes in surface protein expression, remains to be explored. In vitro, such changes are associated with changes in cell phenotype and by inference, changes in the cells' susceptibility to mechanical injury (109, 160, 161). For example, membrane remodeling by loading alveolar epithelial cell membranes with lipids such as cholesterol results in the formation of specialized microdomains (see below), which in turn accelerates transdifferentiation from the Type II to Type I phenotype, which possess different mechanical properties (162, 163). This may also be relevant insofar as the alveolar exudate of injured lungs contains cell debris and is cholesterol rich (164). As the progenitor of the type I cell, during alveolar wound healing, type II cells must divide and differentiate in a cholesterol-rich environment. If and how excess cholesterol effects the differentiation of ATIIs to the ATI phenotype in vivo, and the consequences of this differentiation on deformation-induced lipid trafficking and mechanotransduction, are not known.

    The plasma membrane of eukaryotic cells is enriched in cholesterol and phosphatidylcholine and also contains high levels of sphingolipids (165, 166). Lipids and proteins of the plasma membrane are hydrophobically matched to maintain a low membrane-energy state. Changes in protein and lipid composition alter the membrane energy state and thereby influence cell function (167). Sphingolipids play important roles in a wide variety of cell functions, including mechanotransduction (157). Their concentration in cell membranes is tightly regulated in close association with cholesterol, with which they form membrane microdomains (168, 169). In endothelial cells, fibroblasts, and some epithelial cells, sphingolipids, cholesterol, and GPI-anchored proteins appear to have a preferential association with 50- to 100-nm pits called caveolae as defined by the marker protein caveolin (170). These structures play an important role in noneCclathrin-dependent endocytosis and in contrast to surfactant secreting type II cells can be readily identified in type I alveolar epithelium (171eC174). To the extent to which caveolae are plasma membrane invaginations that may unfold when laterally stressed, they might not only be important for the mechanotransduction of deformation-induced lipid trafficking, but also central for the maintenance of sublytic membrane tension.

    Plasma Membrane Wounding

    Plasma membrane wounding is a common event in exercising muscles and it plays a central role in the pathogenesis of progressive muscle failure in some forms of muscular dystrophy (175, 176). Cell wounding is the reason why patients with increased myocardial stress often have elevated serum levels of "cardiac enzymes," and it probably occurs in the lining cells of the gastrointestinal tract on a regular basis (177, 178). Increasingly, it is being recognized that excessive mechanical forces in the lung result in tissue damage that is characterized by lung cell injury and plasma membrane wounding (24, 95).

    Normally, the plasma membrane carries tensile stress that is at least one order of magnitude lower than that born by filamentous actin (104, 150, 179, 180). The tension at which the plasma membrane fractures is estimated to range between 1 and 25 mN/m, corresponding to membrane strains of only 1 to 3% (181, 182). Lytic tensions vary with the composition and organization of the lipid bilayer as well as with the timeframe for breakage (180, 183). At least in model membranes, lytic tensions are loading rateeCdependent, implying a kinetic process that begins with nucleation of a molecular-scale defect, which either resolves spontaneously or grows to become an unsustainable hole.

    Several studies have examined the molecular as well as biophysical determinants of plasma membrane stress failure in cultured alveolar epithelial cells (108eC112). Results of these studies may be summarized as follows: The probability of stress failure varies with strain amplitude and strain rate. For example, minimal wounding occurs in A549 cells, a human adenocarcinoma cell line, when strain rates are kept at or below 3%/second at a normally injurious strain amplitude (112). The susceptibility for deformation-related stress failure varies considerably between cells and cell culture systems, as does the probability of subsequent membrane repair. Interventions that effect cytoskeletal assembly or vesicular trafficking increase the susceptibility of cells to wounding presumably by impairing their ability to remodel stress-bearing structures (112).

    Plasma Membrane Repair

    The ability to restore membrane integrity after cell wounding is essential for cell survival and virtually all cells possess the means to do so (Figures 3 and 4). Until recently the prevailing view held that injured plasma membranes repaired primarily by "self-sealing" whereby hydrophobic interactions between phospholipids and water would drive lipid flow toward the free edges of a defect (184, 185). Indeed, this mechanism is readily observed in model membranes and red blood cells (186). However, by itself it is insufficient for repair of large wounds and does not account for resealing in nucleated cells (186).

    In 1994, Steinhardt and coworkers described wounding responses in sea urchin eggs and provided the first clues that repair was governed by Ca2+-dependent membrane trafficking and fusion events (187). During the subsequent decade several groups of investigators have extended these observations to mammalian cells and have added considerable detail to our understanding of the responsible molecular mechanisms (reviewed in Reference 175). Small disruptions on the order of 1 e evoke a calcium-dependent exocytosis of vesicles near the wound site, lower plasma membrane tension, and thereby facilitate wound closure (188). The generation and trafficking of vesicles involves nonmuscle myosin and is sensitive to disruptions of the actin cytoskeleton (189, 190). A rise in cytosolic Ca2+ consequent to the loss of plasma membrane integrity also promotes the coalescence of vesicular endomembranes (190eC194). These are transported as a "patch" to the site of larger defects and fuse there with the plasma membrane. Lysosomes appear to be a ubiquitous source of endomembrane patches in wounded cells and the release of lysosomal contents after membrane injury may well be a primitive defense mechanism against invading pathogens (195eC200). Additional novel classes of vesicular organelles have been implicated in cell repair, but their specific roles and functions remain to be defined (201, 202).

    Cells also possess adaptive mechanisms to protect their plasma membranes against repeated mechanical insults. In 3T3 fibroblasts, repeated membrane wounding results in long-term potentiation of Ca2+-regulated vesicular exocytosis in turn generating faster membrane resealing rates (189, 195, 203). This adaptive response requires cAMP-dependent protein kinase A over the short term (minutes) and cAMP response element-binding protein over the long term (days). Vesicular fusion reactions are catalyzed by diverse proteins, which mediate the initial recognition of the membranes that are destined for fusion (196). They pull the membranes close to each other to destabilize the lipid/water interface and to initiate mixing of the lipids. For example, synaptotagmins function as Ca2+ sensors in membrane fusion and play a prominent role in lysosomal exocytosis (204eC206). Synaptotagmin VIIeCdeficient mice show defects in cell resealing and develop a form of autoimmune myositis (207). This suggests that defective membrane repair and the consequent release of intracellular contents overwhelms immune tolerance to self antigens. Although lung morphology and function of synaptotagmin VIIeCdeficient mice have not been characterized to date, it is of note that another syndrome with impaired lysosomal exocytosis, the Hermansky-Pudlak syndrome, is associated with lung pathology (208).

    EFFECTS OF DEFORMING STRESS ON GENE EXPRESSION AND CELL SURVIVAL

    Since the landmark paper by Tremblay and colleagues, which demonstrated a relationship between ventilator settings and inflammatory signaling, the immune and inflammatory responses of lungs to mechanical stress have been extensively studied (209; reviewed in References 13 and 15). Notwithstanding some debate about model, cell, and timing specific differences in the expression of inflammatory mediators, in aggregate the evidence leaves little doubt that inflammation is integral to the pathobiology of the syndrome (210eC213). The concomitant impairment in lung barrier function contributes to the loss of compartmentalization and may account for many of the systemic manifestations of ventilator-associated lung injury (214eC220). The inflammatory effect of mechanical deformation on uninjured lungs ("one hit") compared with preinjured lungs ("two hits") remains an area of important and continued investigation (221).

    The signal transduction pathways that link deformation to some gene response are being characterized in ever-increasing detail (14, 18, 222). Nevertheless, the importance of plasma membrane wounding in initiating a widespread proinflammatory gene response in ventilator-injured lungs is difficult to discern (213, 223). It is clear that not all molecular responses to mechanical ventilation are associated with lung edema or micron-scale plasma membrane lesions (224). Indeed, observations on macrophages and epithelial cells in culture suggest that deforming stress can trigger the release of proinflammatory cytokines in the absence of gross cell injury (211, 212, 225). At the time these observations first appeared, there was some debate if and how an alveolar macrophage might be deformed during mechanical ventilation, and if macrophages or epithelial cells were the predominant source of inflammatory mediators in ventilator-injured lungs. In the interim, the role of epithelial cells as active participants in pulmonary immune responses has been largely acknowledged (226). However, there are virtually no data on epithelial cell strain in situ, let alone data on how an epithelial deformation is transmitted to an adherent macrophage. Be this as it may, many whole animal models of VILI in which inflammatory mechanisms have been characterized have employed ventilation strategies known to produce cell and plasma membrane stress failure (24, 68, 95). When plasma membrane lesions are produced in cell culture, they invariably cause the translocation of nuclear factor-B followed by the induction of early stress response genes (Figure 4) (227). It is reasonable to think that similar events occur in intact mechanically ventilated lungs. Moreover, the resulting induction of CXC chemokines could be amplified and transmitted to uninjured cells by cell contacteC as well as noneCcell contacteCdependent pathways (213, 228eC231). In that scenario, plasma membrane wounding becomes the critical mechanosensing event that initiates and propagates the inflammatory response in the whole organ. Understanding the gene profile of cells that wound, compared with those that are able to prevent or repair wounding, may provide potential protein candidates for further study and chemotherapeutic targeting.

    Cell wounding not only results in gene expression, but can ultimately lead to cell death via cell necrosis or apoptosis. Apoptosis or programmed cell death is an integral mechanism of inflammation, tissue remodeling, and repair. Not surprisingly, therefore, the role of apoptotic mechanisms in lung injury have generated considerable interest, and several reports that injured lungs contain apoptotic cells have appeared (232eC234). While type II cell hyperplasia and cell necrosis dominate the acute phase of VILI, tissue specimens from patients with resolving ARDS have revealed type II cell apoptosis (235). The paucity of apoptotic cells in acutely injured lungs of hyperventilated animals contrasts with the abundance of apoptotic cells in the kidney and GI tract (220). Although it is established that deforming stress can trigger apoptotic responses in lung and muscle cells, it is not known if plasma membrane injury or nuclear deformation are contributing or even necessary priming events (236eC242). In addition, the signals that lead a cell to necrosis or apoptosis remain to be elucidated.

    In the preceding paragraphs, we have highlighted just a few examples of the effects of deforming forces on important cellular responses such as gene expression and resultant protein production, cell proliferation, and survival. Several important issues remain to be explained. First, what is the downstream effect of mechanical forces resulting in cell wounding compared with wounding and subsequent repair What are the determining structural, physiologic, and biological characteristics of cells that fall into each category Finally, what if any difference is there in this cellular response between the cells of the lung (epithelium versus endothelium versus [lipo-]fibroblast)

    CONCLUSIONS

    Motivated by our interest in ventilator-induced and ventilator-associated acute lung injury, we have reviewed the determinants and consequences of the mechanical failure of lung structures. We have focused on tissue elements that are not ordinarily considered important for stress bearing, namely cells and their plasma membranes. Nevertheless, they are susceptible to deformation injury and serve both as sensors and effectors of the innate immune responses that are triggered by physical stress. We have deliberately sought to assimilate observations from nonpulmonary fields to broaden our and the reader's perspective on a problem that should be of interest to practicing intensivists and pulmonary scientists alike.

    Central to our review is the premise that many manifestations of VILI, be they edema, inflammation, or tissue remodeling and fibrosis, can be traced to stress failure of cell membranes. Although injury is not the only trigger of deformation-related cell signaling, one must never forget that it is cells and not alveoli, airways, blood vessels, or connective tissue that sense and transduce local stress. Two interventions, which have received close attention in the critical care literature, namely low tidal volume and high PEEP, clearly reduce cellular stress failure in experimental lung injury models (95, 62). Therefore, one may consider gas exchange, lung aeration, or respiratory mechanics to be surrogates for the real therapeutic objective, namely to prevent cellular stress failure and to enhance airway and alveolar wound healing by physical means. Because wound healing requires cell migration, proliferation, and epithelial transdifferentiation, it is important to understand if and how the aeration of a previously flooded or "closed" airspace influences these critical cell biologic functions. Specifically, there remain fundamental questions about mechanisms through which interfacial forces, cell strain, and gas tension interact to effect epithelial wound repair.

    Intensivists, who interpret gas exchange and mechanics in a cell biological context, may develop a different perspective on rationale and efficacy of certain treatment approaches. For example, some believe that bilevel pressure ventilation obviates the need for limiting tidal volume in mechanically ventilated patients with injured lungs (243). This opinion is largely grounded in the observation that preserving diaphragm activity prevents the atelectasis of diaphragm apposed dependent lung. In other words, the loss of regional aeration is equated with injury. However, the prognostic relevance of the surrogate endpoint, atelectasis, remains unclear. Cells are injured because the matrix to which they adhere undergoes large deformations or because their apical membranes are "abraded" by the cyclic movement of aireCliquid interfaces and foam across them (64, 112). None of the observed effects of bilevel pressure ventilation on regional lung aeration directly address these cell injury mechanisms. Moreover, the sometimes overlooked work by Mascheroni and colleagues argues that hyperventilation-induced lung injury need not be restricted to positive pressure breathing (244). This is because large oscillations in alveolar volume and surface area, be they generated by machines or the respiratory muscles, over time impair the physicochemical properties of surfactant (245eC247). The point of this example is not to refute claims of efficacy of bilevel pressure ventilation, but only to raise caution against the ready acceptance of a surrogate physiologic endpoint as proof of benefit.

    The focus on cellular stress failure as central to VILI also suggests new treatment targets. These include molecules involved in the regulation of deformation-induced cytoskeletal remodeling, lipid trafficking, and membrane repair. Much preclinical work will need to be done before related approaches can be tested at the bedside. On the other hand, some "old drugs" that are currently in use for different indications may already point in this direction. For example, it has been known for some time that -adrenergic receptor agonists preserve barrier properties of ventilator-injured lungs (248). Among the many putative mechanisms of benefit is a cyclic adenosine monophosphate (cAMP)-mediated effect on the endothelial cell cytoskeleton. To think that a change in the resistance of endothelial cells to deforming stress could alter their susceptibility for stress failure would take but a small leap of faith and is an intriguing and testable hypothesis. The current approach to mechanical ventilation is built on the foundations of classic cardiopulmonary physiology and respiratory system mechanics. In dealing with VILI, the critical care community has in the past decade discovered an important connection between mechanics and innate immunity. The term "biotrauma" coined by Arthur Slutsky embraces and underscores this connection (249). Has the time come to add plasma membrane and cytoskeletal biology to the topics an intensivist should know something about We hope that this review convinces the reader that the answer is "yes."

    REFERENCES

    Rubenfeld GD. Epidemiology of acute lung injury. Crit Care Med 2003;31:S276eCS284.

    Moss M, Mannino DM. Race and gender differences in acute respiratory distress syndrome deaths in the United States: An analysis of multiple cause mortality data (1979eC1996). Crit Care Med 2002;30:1679eC1685.

    Bersten AD, Edibam C, Hunt T, Moran J. Incidence and mortality of acute lung injury and the acute respiratory distress syndrome in three Australian States. Am J Respir Crit Care Med 2002;165:443eC448.

    Milberg JA, Davis DR, Steinberg KP, Hudson LD. Improved survival of patients with acute respiratory distress syndrome (ARDS): 1983eC1993. JAMA 1995;273:306eC309.

    Luhr OR, Antonsen K, Karlsson M, Aardal S, Thorsteinsson A, Frosell CG, Bonde J. Incidence and mortality after acute respiratory failure and acute respiratory distress syndrome in Sweden, Denmark, and Iceland. The ARF Study Group. Am J Respir Crit Care Med 1999;159:1849eC1861.

    Abel SJ, Finney SJ, Brett SJ, Keogh BF, Morgan CJ, Evans TW. Reduced mortality in association with the acute respiratory distress syndrome (ARDS). Thorax 1998;53:292eC294.

    The Acute Respiratory Distress Syndrome Network. Ventilation with lower tidal volumes as compared with traditional tidal volumes for acute lung injury and the acute respiratory distress syndrome. N Engl J Med 2000;342:1301eC1308.

    Amato MBP, Barbas CSV, Medeiros DM, Magaldi RB, Schettino GDP, Lorenzi G, Kairalla RA, Deheinzelin D, Munoz C, Oliveira R, et al. Effect of a protective-ventilation strategy on mortality in the acute respiratory distress syndrome. N Engl J Med 1998;338:347eC354.

    Dreyfuss D, Saumon G. Ventilator-induced lung injury: lessons from experimental studies. Am J Respir Crit Care Med 1998;157:294eC323.

    Whitehead T, Slutsky AS. The pulmonary physician in critical care: 7. Ventilator-induced lung injury. Thorax 2002;7:635eC642.

    Uhlig S. Mechanotransduction in the lung: ventilation-induced lung injury and mechanotransduction: stretching it too far Am J Physiol Lung Cell Mol Physiol 2002;282:L892eCL896.

    Frank JA, Matthay MA. Science review: mechanisms of ventilator-induced injury. Crit Care 2003;7:233eC241.

    Pugin J. Molecular mechanisms of lung cell activation induced by cyclic stretch. Crit Care Med 2003;31:S200eCS206.

    Liu M, Tanswell AK, Post M. Mechanical force-induced signal transduction in lung cells. Am J Physiol 1999;277:L667eCL683.

    Dos Santos CC, Slutsky AS. Invited review: mechanisms of ventilator-induced lung injury: a perspective. J Appl Physiol 2000;89:1645eC1655.

    Wirtz HR, Dobbs LG. The effects of mechanical forces on lung functions. Respir Physiol 2000;119:1eC17.

    Vlahakis NE, Hubmayr RD. Response of alveolar cells to mechanical stress. Curr Opin Crit Care 2003;9:2eC8.

    Shyy JY, Chien S. Role of integrins in endothelial mechanosensing of shear stress. Circ Res 2002;91:769eC775.

    Edwards YS. Stretch stimulation: its effects on alveolar type II cell function in the lung. Comp Biochem Physiol A Mol Integr Physiol 2001;129:245eC260.

    Savov J, Silbajoris R, Young SL. Mechanical ventilation of rat lung: effect on surfactant forms. Am J Physiol 1999;277:L320eCL326.

    Malloy JL, Veldhuizen RA, Lewis JF. Effects of ventilation on the surfactant system in sepsis-induced lung injury. J Appl Physiol 2000;88:401eC408.

    Torday JS, Rehan VK. Stretch-stimulated surfactant synthesis is coordinated by the paracrine actions of PTHrP and leptin. Am J Physiol Lung Cell Mol Physiol 2002;283:L130eCL135.

    American Thoracic Society/European Respiratory Society. International consensus conferences in intensive care medicine: ventilator-associated lung injury in ARDS. Am J Respir Crit Care Med 1999;160:2118eC2124.

    Dreyfuss D, Basset G, Soler P, Saumon G. Intermittent positive-pressure hyperventilation with high inflation pressures produces pulmonary microvascular injury in rats. Am Rev Respir Dis 1985;132:880eC884.

    Fu Z, Costello ML, Tsukimoto K, Prediletto R, Elliott AR, Mathieu-Costello O, West JB. High lung volumes increases stress failure in pulmonary capllaries. J Appl Physiol 1992;73:123eC133.

    Costello ML, Mathieu-Costello O, West JB. Stress failure of alveolar epithelial cells studied by scanning electron microscopy. Am Rev Respir Dis 1992;145:1446eC1455.

    John E, McDevitt M, Wilborn W, Cassady G. Ultrastructure of the lung after ventilation. Br J Exp Pathol 1982;63:401eC407.

    Parker JC, Townsley MI. Evaluation of lung injury in rats and mice. Am J Physiol Lung Cell Mol Physiol 2004;286:L231eCL246.

    Agostoni E. Mechanics of the pleural space. In: Geiger SR, editor. Handbook of physiology. Section 3: The respiratory system. Bethesda, MD: American Physiological Society; 1986. pp. 531eC559.

    D'Angelo E, Michelini S, Agostoni E. Partition of factor contributing to the vertical gradient of transpulmonary pressure. Respir Physiol 1971;12:90eC101.

    Wilson TA. Solid mechanics. In: Fishman AP, editor. Handbook of Physiology. Section 3: Respiratory System, Volume III: Mechanics of Breathing, Part I. Baltimore, MD: Williams and Wilkins Co.; 1986. pp. 35eC40.

    Rodarte JR, Fung YC. Distribution of stresses within the lung. In: Fishman AP, editor. Handbook of Physiology. Section 3: Respiratory System. Volume III: Mechanics of Breathing, Part I. Baltimore, MD: Williams and Wilkins Co.; 1986. pp. 233eC246.

    Bar-Yishay E, Hyatt RE, Rodarte JR. Effect of heart weight on distribution of lung surface pressures in vertical dogs. J Appl Physiol 1986;61:712eC718.

    Agostoni E, D'Angelo E, Bonanni MV. The effect of the abdomen on the vertical gradient of pleural surface pressure. Respir Physiol 1970;8:332eC346.

    Chang H, Lai-Fook SJ, Domino KB, Schimmel C, Hildebrandt J, Robertson HT, Glenny RW, Hlastala MP. Spatial distribution of ventilation and perfusion in anesthetized dogs in lateral postures. J Appl Physiol 2002;92:745eC762.

    Hubmayr RD, Rodarte JR, Walters BJ, Tonelli FM. Regional ventilation during spontaneous breathing and mechanical ventilation in dogs. J Appl Physiol 1987;63:2467eC2475.

    Rodarte JR, Hubmayr RD, Stamenovic D, Walters BJ. Regional lung strain in dogs during deflation from total lung capacity. J Appl Physiol 1985;58:164eC172.

    Bachofen H, Schurch S. Alveolar surface forces and lung architecture. Comp Biochem Physiol A Mol Integr Physiol 2001;129:183eC193.

    Lai-Fook SJ, Kallok MJ. Bronchial-arterial interdependence in isolated dog lung. J Appl Physiol 1982;52:1000eC1007.

    Bachofen H, Schurch S, Urbinelli M, Weibel ER. Relations among alveolar surface tension, surface area, volume, and recoil pressure. J Appl Physiol 1987;62:1878eC1887.

    Mercer RR, Laco JM, Crapo JD. Three-dimensional reconstruction of alveoli in the rat lung for pressure-volume relationships. J Appl Physiol 1987;62:1480eC1487.

    Mercer RR, Crapo JD. Spatial distribution of collagen and elastin fibers in the lungs. J Appl Physiol 1990;69:756eC765.

    Gil J, Bachofen H, Gehr P, Weibel ER. Alveolar volume-surface area relation in air- and saline-filled lungs fixed by vascular perfusion. J Appl Physiol 1979;47:990eC1001.

    Oldmixon EH, Hoppin FG Jr. Distribution of elastin and collagen in canine lung alveolar parenchyma. J Appl Physiol 1989;67:1941eC1949.

    Oldmixon EH, Hoppin FG Jr. Alveolar septal folding and lung inflation history. J Appl Physiol 1991;71:2369eC2379.

    Wilson TA, Bachofen H. A model for mechanical structure of the alveolar duct. J Appl Physiol 1982;52:1064eC1070.

    Tschumperlin DJ, Margulies SS. Alveolar epithelial surface area-volume relationship in isolated rat lungs. J Appl Physiol 1999;86:2026eC2033.

    Bachofen H, Gerber U, Schurch S. Effects of fixatives on function of pulmonary surfactant. J Appl Physiol 2002;93:911eC916.

    Carney DE, Bredenberg CE, Schiller HJ, Picone AL, McCann UG, Gatto LA, Bailey G, Fillinger M, Nieman GF. The mechanism of lung volume change during mechanical ventilation. Am J Respir Crit Care Med 1999;160:1697eC1702.

    Halter JM, Steinberg JM, Schiller HJ, DaSilva M, Gatto LA, Lanclas S, Nieman GF. Positive end-expiratory pressure after a recruitment maneuver prevents both alveolar collapse and recruitment/derecruitment. Am J Respir Crit Care Med 2003;167:1620eC1626.

    Schiller HJ, McCann UG II, Carney DE, Gatto LA, Steinberg JM, Nieman GF. Altered alveolar mechanics in the acutely injured lung. Crit Care Med 2001;29:1049eC1055.

    Steinberg J, Schiller HJ, Halter JM, Gatto LA, Dasilva M, Amato M. MCCann UG, Nieman GF. Tidal volume increases do not affect alveolar mechanics in normal lung but cause alveolar overdistension and exacerbate alveolar instability after surfactant deactivation. Crit Care Med 2002;30:2675eC2683.

    Bachofen H, Schech S, Michel RP, Weibel ER. Experimental hydrostatic pulmonary edema in rabbit lungs: Morphology. Am Rev Respir Dis 1993;147:989eC996.

    Bachofen H, Schech S, Weibel ER. Experimental hydrostatic pulmonary edema in rabbit lungs: barrier lesions. Am Rev Respir Dis 1993;147:997eC1004.

    Fehrenbach A, Fehrenbach H, Wittwer T, Ochs M, Wahlers T, Richter J. Evaluation of pulmonary edema: stereological versus gravimetrical analysis. Eur Surg Res 2001;33:270eC278.

    Hubmayr RD. Perspective on lung injury and recruitment: a skeptical look at the opening and collapse story. Am J Respir Crit Care Med 2002;165:1647eC1653.

    Wilson TA, Anafi RC, Hubmayr RD. Mechanics of edematous lungs. J Appl Physiol 2001;90:2088eC2093.

    Martynowicz MA, Minor TA, Walters BJ, Hubmayr RD. Regional expansion of oleic acid-injured lungs. Am J Respir Crit Care Med 1999;160:250eC258.

    Maunder RJ, Shuman WP, McHugh JW, Marglin SI, Butler J. Preservation of normal lung regions in the adult respiratory distress syndrome: analysis by computed tomography. JAMA 1986;255:2463eC2465.

    Gattinoni L, Pesenti A, Avalli L, Rossi F, Bombino M. Pressure-volume curve of total respiratory system in acute respiratory failure: computed tomographic scan study. Am Rev Respir Dis 1987;136:730eC736.

    Mead J, Takishima T, Leith D. Stress distribution in lungs: a model of pulmonary elasticity. J Appl Physiol 1970;28:596eC608.

    Muscedere JG, Mullen JB, Gan K, Slutsky AS. Tidal ventilation at low airway pressures can augment lung injury. Am J Respir Crit Care Med 1994;149:1327eC1334.

    Goldstein I, Bughalo MT, Marquette CH, Lenaour G, Lu Q, Rouby JJ. Mechanical ventilation-induced air-space enlargement during experimental pneumonia in piglets. Am J Respir Crit Care Med 2001;163:958eC964.

    Bilek AM, Dee KC, Gaver DP III. Mechanisms of surface-tension-induced epithelial cell damage in a model of pulmonary airway reopening. J Appl Physiol 2003;94:770eC783.

    Gaver DP III, Kute SM. A theoretical model study of the influence of fluid stresses on a cell adhering to a microchannel wall. Biophys J 1998;75:721eC733.

    Cassidy KJ, Gavriely N, Grotberg JB. Liquid plug flow in straight and bifurcating tubes. J Biomech Eng 2001;123:580eC589.

    Greenfield LJ, Ebert PA, Benson DW. Effect of positive pressure ventilation on surface tension of lung extract. Anesthesiology 1964;25:312eC316.

    Webb HH, Tierney DF. Experimental pulmonary edema due to intermittent positive pressure ventilation with high inflation pressures: protection by positive end-expiratory pressure. Am Rev Respir Dis 1974;110:556eC565.

    Egan EA. Lung inflation, lung solute permeability and alveolar edema. J Appl Physiol 1982;53:121eC125.

    Parker JC, Townsley MI, Rippe B, Taylor AE, Thigpen J. Increased microvascular permeability in dog lungs due to high peak airway pressures. J Appl Physiol 1984;57:1809eC1816.

    Parker JC, Breen EC, West JB. High vascular and airway pressures increase interstitial protein mRNA expression in isolated rat lungs. J Appl Physiol 1997;83:1697eC1705.

    Parker JC, Ivey CL, Tucker A. Phosphotyrosine phosphatase and tyrosine kinase inhibition modulate airway pressure-induced lung injury. J Appl Physiol 1998;85:1753eC1761.

    Parker JC, Ivey CL, Tucker JA. Gadolinium prevents high airway pressure-induced permeability increases in isolated rat lungs. J Appl Physiol 1998;84:1113eC1118.

    West JB. Thoughts on the pulmonary blood-gas barrier. EB2003 Comroe Lecture. Am J Physiol Lung Cell Mol Physiol 2003;285:L501eCL513.

    West JB, Tsukimoto K, Mathieu-Costello O, Prediletto R. Stress failure in pulmonary capillaries. J Appl Physiol 1991;70:1731eC1742.

    Tsukimoto K, Mathieu-Costello O, Prediletto R, Elliott AR, West JB. Ultrastructural appearances of pulmonary capillaries at high transmural pressures. J Appl Physiol 1991;71:573eC582.

    Swenson ER, Maggiorini M, Mongovin S, Gibbs JS, Greve I, Mairbaurl H, Bartsch P. Pathogenesis of high altitude pulmonary edema: inflammation is not an etiologic factor. JAMA 2002;287:2228eC2235.

    Wieslander J, Heinegard D. The involvement of type IV collagen in Goodpasture's Syndrome. Ann N Y Acad Sci 1985;460:363eC374.

    Hopkins SR, Schoene RB, Martin TR, Henderson WR, Spragg RG, West JB. Intense exercise impairs the integrity of the pulmonary blood gas barrier in elite athletes. Am J Respir Crit Care Med 1997;155:1090eC1094.

    Hotchkiss JR, Simonson DA, Marek DJ, Marini JJ, Dries DJ. Pulmonary microvascular fracture in a patient with acute respiratory distress syndrome. Crit Care Med 2002;30:2368eC2370.

    Broccard A, Shapiro RS, Schmitz LL, Adams AB, Nahum A, Marini JJ. Prone positioning attenuates and redistributes ventilator-induced lung injury in dogs. Crit Care Med 2000;28:295eC303.

    Majno G, Shea SM, Leventhal M. Endothelial contraction induced by histamine-type mediators: an electron microscopic study. J Cell Biol 1969;42:647eC672.

    Dudek SM, Garcia JG. Cytoskeletal regulation of pulmonary vascular permeability. J Appl Physiol 2001;91:1487eC1500.

    Bogatcheva NV, Dudek SM, Garcia JG, Verin AD. Mitogen-activated protein kinases in endothelial pathophysiology. J Investig Med 2003;51:341eC352.

    Bogatcheva N, Garcia JG, Verin AD. Role of tyrosine kinase signaling in endothelial cell barrier regulation. Vascul Pharmacol 2002;39:201eC212.

    Hammersen F, Hammersen E. The ultrastructure of endothelial gap formation and leukocyte emigration. In: Messmer K, Hammersen F, editors. Microcirculation and inflammation: vessel wall, inflammatory cells, mediator interation. Proceedings of the 6th Bodensee Symposium on Microcirculation 1986 Jun 22eC25; Heidelberg, Germany. Basel: Karger, 1987. p. 1eC34.

    Bhattacharya S, Sen N, Yiming MT, Patel R, Parthasarathi K, Quadri S, Issekutz AC, Bhattacharya J. High tidal volume ventilation induces proinflammatory signaling in rat lung endothelium. Am J Respir Cell Mol Biol 2003;28:218eC224.

    Neal CR, Michel CC. Transcellular openings through microvascular walls in acutely inflamed frog mesentery. Exp Physiol 1992;77:917eC920.

    Neal CR, Michel CC. Transcellular gaps in microvascular walls of frog and rat when permeability is increased by perfusion with the ionophore A23187. J Physiol 1995;488:427eC437.

    Neal CR, Michel CC. Openings in frog microvascular endothelium induced by high intravascular pressures. J Physiol 1996;492:39eC52.

    Feng D, Nagy NA, Hipp J, Pyne K, Dvorak HF, Dvorak AM. Reinterpretation of endothelial cell gaps induced by vasoactive mediators in guinea-pig, mouse and rat: many are transcellular pores. J Physiol 1997;504:747eC761.

    Neal CR, Michel CC. Transcellular openings through frog microvascular endothelium. Exp Physiol 1997;82:419eC422.

    Elliott AR, Fu Z, Tsukimoto K, Prediletto R, Mathieu-Costello O, West JB. Short-term reversibility of ultrastructural changes in pulmonary capillaries caused by stress failure. J Appl Physiol 1992;73:1150eC1158.

    Dreyfuss D, Soler P, Saumon G. Spontaneous resolution of pulmonary edema caused by short periods of cyclic overinflation. J Appl Physiol 1992;72:2081eC2089.

    Gajic O, Lee J, Doerr CH, Berrios JC, Myers JL, Hubmayr RD. Ventilator-induced cell wounding and repair in the intact lung. Am J Respir Crit Care Med 2003;167:1057eC1063.

    Neal CR, Michel CC. Effects of temperature on the wall strength and compliance of frog mesenteric microvessels. J Physiol 2000;526:613eC622.

    Parker JC, Miller GT, Tarpey SB, Anghelescu M, Penton AM, Adkison JB. Cell wounding modulates junctional proteins in rat pulmonary artery and microvascular endothelial cells . Am J Respir Crit Care Med 2004;169:A158.

    Chen CS, Tan J, Tien J. Mechanotransduction at cell-matrix and cell-cell contacts. Annu Rev Biomed Eng 2004;6:275eC302.

    Huang H, Kamm RD, Lee RT. Cell mechanics and mechanotransduction: pathways, probes, and physiology. Am J Physiol Cell Physiol 2004;287:C1eC11.

    Janmey PA, Weitz DA. Dealing with mechanics: mechanisms of force transduction in cells. Trends Biochem Sci 2004;29:364eC370.

    Geiger B, Bershadsky A, Pankov R, Yamada KM. Transmembrane crosstalk between extracellular matrix-cytoskeleton crosstalk. Nat Rev Mol Cell Biol 2001;2:793eC805.

    Hynes RO. Integrins: bidirectional, allosteric signaling machines. Cell 2002;110:673eC687.

    Ingber DE. Opposing views on tensegrity as a structural framework for understanding cell mechanics. J Appl Physiol 2000;89:1663eC1670.

    Wang N, Naruse K, Stamenovic D, Fredberg JJ, Mijailovich SM, Tolic-Norrelykke IM, Polte T, Mannix R, Ingber DE. Mechanical behavior in living cells consistent with the tensegrity model. Proc Natl Acad Sci USA 2001;98:7765eC7770.

    Schwarz US, Balaban NQ, Riveline D, Addadi L, Bershadsky A, Safran SA, Geiger B. Measurement of cellular forces at focal adhesions using elastic micro-patterned substrates. Mater Sci Eng C-Biomimetic Supramol Syst 2003;23:387eC394.

    Parton RG, Hancock JF. Lipid rafts and plasma membrane microorganization: insights from Ras. Trends Cell Biol 2004;14:141eC147.

    Stamenovic D, Wang N. Invited review: engineering approaches to cytoskeletal mechanics. J Appl Physiol 2000;89:2085eC2090.

    Oswari J, Matthay MA, Margulies SS. Keratinocyte growth factor reduces alveolar epithelial susceptibility to in vitro mechanical deformation. Am J Physiol Lung Cell Mol Physiol 2001;281:L1068eCL1077.

    Tschumperlin DJ, Margulies SS. Equibiaxial deformation-induced injury of alveolar epithelial cells in vitro. Am J Physiol 1998;275:L1173eCL1183.

    Stroetz RW, Vlahakis NE, Walters BJ, Schroeder MA, Hubmayr RD. Validation of a new live cell strain system: characterization of plasma membrane stress failure. J Appl Physiol 2001;90:2361eC2370.

    Tschumperlin DJ, Oswari J, Margulies AS. Deformation-induced injury of alveolar epithelial cells. Effect of frequency, duration, and amplitude. Am J Respir Crit Care Med 2000;162:357eC362.

    Vlahakis NE, Schroeder MA, Pagano RE, Hubmayr RD. Role of deformation-induced lipid trafficking in the prevention of plasma membrane stress failure. Am J Respir Crit Care Med 2002;166:1282eC1289.

    Berrios JC, Schroeder MA, Hubmayr RD. Mechanical properties of alveolar epithelial cells in culture. J Appl Physiol 2001;91:65eC73.

    Doornaert B, Leblond V, Planus E, Galiacy S, Laurent VM, Gras G, Isabey D, Lafuma C. Time course of actin cytoskeleton stiffness and matrix adhesion molecules in human bronchial epithelial cell cultures. Exp Cell Res 2003;287:199eC208.

    Coughlin MF, Stamenovic D. A prestressed cable network model of the adherent cell cytoskeleton. Biophys J 2003;84:1328eC1336.

    Fabry B, Maksym GN, Butler JP, Glogauer M, Navajas D, Fredberg JJ. Scaling the microrheology of living cells. Phys Rev Lett 2001;87:148102.

    Ingber DE, Tensegrity I. Cell structure and hierarchical systems biology. J Cell Sci 2003;116:1157eC1173.

    Gunst SJ, Tang DD, Saez AO. Cytoskeletal remodeling of the airway smooth muscle cell: a mechanism for adaptation to mechanical forces in the lung. Respir Physiol Neurobiol 2003;137:151eC168.

    Stossel TP, Hartwig JH. Filling gaps in signaling to actin cytoskeletal remodeling. Dev Cell 2003;4:444eC445.

    Shyy JY, Chien S. Role of integrins in endothelial mechanosensing of shear stress. Circ Res 2002;91:769eC775.

    Yoshigi M, Clark EB, Yost HJ. Quantification of stretch-induced cytoskeletal remodeling in vascular endothelial cells by image processing. Cytometry 2003;55A:109eC118.

    Wang N, Ingber DE. Control of cytoskeletal mechanics by extracellular matrix, cell shape, and mechanical tension. Biophys J 1994;66:2181eC2189.

    Heidemann SR, Wirtz D. Towards a regional approach to cell mechanics. Trends Cell Biol 2004;14:160eC166.

    Ko KS, McCulloch CA. Partners in protection: interdependence of cytoskeleton and plasma membrane in adaptions to applied forces. J Membr Biol 2000;174:85eC95.

    Vlahakis NE, Schroeder MA, Pagano RE, Hubmayr RD. Deformation-induced lipid trafficking in alveolar epithelial cells. Am J Physiol Lung Cell Mol Physiol 2001;280:L938eCL946.

    Bischofs IB, Schwarz US. Cell organization in soft media due to active mechanosensing. Proc Natl Acad Sci USA 2003;100:9274eC9279.

    Bershadsky AD, Balaban NQ, Geiger B. Adhesion-dependent cell mechanosensitivity. Annu Rev Cell Dev Biol 2003;19:677eC695.

    Shafrir Y, Forgacs G. Mechanotransduction through the cytoskeleton. Am J Physiol Cell Physiol 2002;282:C479eCC486.

    Janmey PA. The cytoskeleton and cell signaling: component localization and mechanical coupling. Physiol Rev 1998;78:763eC781.

    Jalali S, del Pozo MA, Chen K, Miao H, Li Y, Schwartz MA, Shyy JY, Chien S. Integrin-mediated mechanotransduction requires its dynamic interaction with specific extracellular matrix (ECM) ligands. Proc Natl Acad Sci USA 2001;98:1042eC1046.

    Li S, Chen BP, Azuma N, Hu YL, Wu SZ, Sumpio BE, Shyy JY, Chien S. Distinct roles for the small GTPases Cdc42 and Rho in endothelial responses to shear stress. J Clin Invest 1999;103:1141eC1150.

    Schmidt C, Pommerenke H, Durr F, Nebe B, Rychly J. Mechanical stressing of integrin receptors induces enhanced tyrosine phosphorylation of cytoskeletally anchored proteins. J Biol Chem 1998;273:5081eC5085.

    Juliano RL. Signal transduction by cell adhesion receptors and the cytoskeleton: functions of integrins, cadherins, selectins, and immunoglobulin-superfamily members. Annu Rev Pharmacol Toxicol 2002;42:283eC323.

    Li S, Kim M, Hu YL, Jalali S, Schlaepfer DD, Hunter T, Chien S, Shyy JY. Fluid shear stress activation of focal adhesion kinase: linking to mitogen-activated protein kinases. J Biol Chem 1997;272:30455eC30462.

    Chen KD, Li YS, Kim M, Li S, Yuan S, Chien S, Shyy JY. Mechanotransduction in response to shear stress. Roles of receptor tyrosine kinases, integrins, and Shc. J Biol Chem 1999;274:18393eC18400.

    Wang Y, Miao H, Li S, Chen KD, Li YS, Yuan S, Shyy JY, Chien S. Interplay between integrins and FLK-1 in shear stress-induced signaling. Am J Physiol Cell Physiol 2002;283:C1540eCC1547.

    Takenawa T, Itoh T. Phosphoinositides, key molecules for regulation of actin cytoskeletal organization and membrane traffic from the plasma membrane. Biochim Biophys Acta 2001;1533:190eC206.

    Go YM, Park H, Maland MC, Darley-Usmar VM, Stoyanov B, Wetzker R, Jo H. Phosphatidylinositol 3-kinase gamma mediates shear stress-dependent activation of JNK in endothelial cells. Am J Physiol 1998;275:H1898eCH1904.

    Cavanaugh KJ Jr, Oswari J, Margulies SS. Role of stretch on tight junction structure in alveolar epithelial cells. Am J Respir Cell Mol Biol 2001;25:584eC591.

    Cavanaugh KJ Jr, Margulies SS. Measurement of stretch-induced loss of alveolar epithelial barrier integrity with a novel in vitro method. Am J Physiol Cell Physiol 2002;283:C1801eCC1808.

    Berg JT, Fu Z, Breen EC, Tran HC, Mathieu-Costello O, West JB. High lung inflation increases mRNA levels of ECM components and growth factors in lung parenchyma. J Appl Physiol 1997;83:120eC128.

    Sheppard D. Functions of pulmonary epithelial integrins: from development to disease. Physiol Rev 2003;83:673eC686.

    Steinberg J, Halter J, Schiller HJ, Dasilva M, Landas S, Gatto LA, Maisi P, Sorsa T, Rajamaki M, Lee HM, et al. Metalloproteinase inhibition reduces lung injury and improves survival after cecal ligation and puncture in rats. J Surg Res 2003;111:185eC195.

    Swartz MA, Tschumperlin DJ, Kamm RD, Drazen JM. Mechanical stress is communicated between different cell types to elicit matrix remodeling. Proc Natl Acad Sci USA 2001;98:6180eC6185.

    Tschumperlin DJ, Dai G, Maly IV, Kikuchi T, Laiho LH, McVittie AK, Haley KJ, Lilly CM, So PT, Lauffenburger DA, et al. Mechanotransduction through growth-factor shedding into the extracellular space. Nature 2004;429:83eC86.

    Foda HD, Rollo EE, Drews M, Conner C, Appelt K, Shalinsky DR, Zucker S. Ventilator-induced lung injury upregulates and activates gelatinases and EMMPRIN: attenuation by the synthetic matrix metalloproteinase inhibitor, Prinomastat (AG3340). Am J Respir Cell Mol Biol 2001;25:717eC724.

    Haseneen NA, Vaday GG, Zucker S, Foda HD. Mechanical stretch induces MMP-2 release and activation in lung endothelium: role of EMMPRIN. Am J Physiol Lung Cell Mol Physiol 2003;284:L541eCL547.

    Fisher JL, Levitan I, Margulies SS. Plasma membrane surface increases with tonic stretch of alveolar epithelial cells. Am J Respir Cell Mol Biol 2004;31:200eC208.

    Morris CE. How did cells get their size Anat Rec 2002;268:239eC251.

    Sheetz MP. Cell control by membrane-cytoskeleton adhesion. Nat Rev Mol Cell Biol 2001;2:392eC396.

    Morris CE, Homann U. Cell surface area regulation and membrane tension. J Membr Biol 2001;179:79eC102.

    Raucher D, Sheetz MP. Characteristics of a membrane reservoir buffering membrane tension. Biophys J 1999;77:1992eC2002.

    Wirtz HR, Dobbs LG. Calcium mobilization and exocytosis after one mechanical stretch of lung epithelial cells. Science 1990;250:1266eC1269.

    Frick M, Bertocchi C, Jennings P, Haller T, Mair N, Singer W, Pfaller W, Ritsch-Marte M, Dietl P. Ca2+ entry is essential for cell strain-induced lamellar body fusion in isolated rat type II pneumocytes. Am J Physiol Lung Cell Mol Physiol 2004;286:L210eCL220.

    Truschel ST, Wang E, Ruiz WG, Leung SM, Rojas R, Lavelle J, Zeidel M, Stoffer D, Apodaca G. Stretch-regulated exocytosis/endocytosis in bladder umbrella cells. Mol Biol Cell 2002;13:830eC846.

    Berrios JC, Hubmayr RD. Deforming stress triggers endocytosis in alveolar epithelial cells . Am J Respir Crit Care Med 2003;167:A57.

    Volonte D, Galbiati F, Pestell RG, Lisanti MP. Cellular stress induces the tyrosine phosphorylation of caveolin-1 (Tyr(14)) via activation of p38 mitogen-activated protein kinase and c-Src kinase: evidence for caveolae, the actin cytoskeleton, and focal adhesions as mechanical sensors of osmotic stress. J Biol Chem 2001;276:8094eC8103.

    Palestini P, Calvi C, Conforti E, Daffara R, Botto L, Miserocchi G. Compositional changes in lipid microdomains of air-blood barrier plasma membranes in pulmonary interstitial edema. J Appl Physiol 2003;95:1446eC1452.

    Daffara R, Botto L, Beretta E, Conforti E, Faini A, Palestini P, Miserocchi G. Endothelial cells as early sensors of pulmonary interstitial edema. J Appl Physiol 2004;97:1575eC1583.

    Gutierrez JA, Ertsey R, Scavo LM, Collins E, Dobbs LG. Mechanical distention modulates alveolar epithelial cell phenotypic expression by transcriptional regulation. Am J Respir Cell Mol Biol 1999;21:223eC229.

    Gutierrez JA, Suzara VV, Dobbs LG. Continuous mechanical contraction modulates expression of alveolar epithelial cell phenotype. Am J Respir Cell Mol Biol 2003;29:81eC87.

    Kolleck I, Guthmann F, Ladhoff AM, Tandon NN, Schlame M, Rustow B. Cellular cholesterol stimulates acute uptake of palmitate by redistribution of fatty acid translocase in type II pneumocytes. Biochemistry 2002;41:6369eC6375.

    Kolleck I, Wissel H, Guthmann F, Schlame M, Sinha P, Rustow B. HDL-holoparticle uptake by alveolar type II cells: effect of vitamin E status. Am J Respir Cell Mol Biol 2002;27:57eC63.

    Swendsen CL, Skita V, Thrall RS. Alterations in surfactant neutral lipid composition during the development of bleomycin-induced pulmonary fibrosis. Biochim Biophys Acta 1996;1301:90eC96.

    Brown DA, London E. Structure and function of sphingolipid- and cholesterol-rich membrane rafts. J Biol Chem 2000;275:17221eC17224.

    Pagano RE, Watanabe R, Wheatley C, Dominguez M. Applications of BODIPY-sphingolipid analogs to study lipid traffic and metabolism in cells. Methods Enzymol 2000;312:523eC534.

    Mitra K, Ubarretxena-Belandia I, Taguchi T, Warren G, Engelman DM. Modulation of the bilayer thickness of exocytic pathway membranes by membrane proteins rather than cholesterol. Proc Natl Acad Sci USA 2004;101:4083eC4088.

    Simons K, Toomre D. Lipid rafts and signal transduction. Nat Rev Mol Cell Biol 2000;1:31eC39.

    van Meer G, Holthuis JCM. Sphingolipid transport in eukaryotic cells. Biochim Biophys Acta 2000;1486:145eC170.

    Razani B, Woodman SE, Lisanti MP. Caveolae: from cell biology to animal physiology. Pharmacol Rev 2002;54:431eC467.

    Ludger J, Lamaze C. Clathrin-dependent or not: is it still the question Traffic 2002;3:443eC445.

    Gil J. Number and distribution of plasmalemmal vesicles in the lung. Fed Proc 1983;42:2414eC2418.

    Kasper M, Reimann T, Hempel U, Wenzel KW, Bierhaus A, Schuh D, Dimmer V, Haroske G, Muller M. Loss of caveolin expression in type I pneumocytes as an indicator of subcellular alterations during lung fibrogenesis. Histochem Cell Biol 1998;109:41eC48.

    Newman GR, Campbell L, von Ruhland C, Jasani B, Gumbleton M. Caveolin and its cellular and subcellular immunolocalisation in lung alveolar epithelium: implications for alveolar epithelial type I cell function. Cell Tissue Res 1999;295:111eC120.

    McNeil PL, Steinhardt RA. Plasma membrane disruption: repair, prevention, adaptation. Annu Rev Cell Dev Biol 2003;19:697eC731.

    Bansal D, Miyake K, Vogel SS, Groh S, Chen CC, Williamson R, McNeil PL, Campbell KP. Defective membrane repair in dysferlin-deficient muscular dystrophy. Nature 2003;423:168eC172.

    Fischer TA, McNeil PL, Khakee R, Finn P, Kelly RA, Pfeffer MA, Pfeffer JM. Cardiac myocyte membrane wounding in the abruptly pressure-overloaded rat heart under high wall stress. Hypertension 1997;30:1041eC1046.

    McNeil PL, Ito S. Gastrointestinal cell plasma membrane wounding and resealing in vivo. Gastroenterology 1989;96:1238eC1248.

    Olbrich K, Rawicz W, Needham D, Evans E. Water permeability and mechanical strength of polyunsaturated lipid bilayers. Biophys J 2000;79:321eC327.

    Waugh RE. Effects of abnormal cytoskeletal structure on erythrocyte membrane mechanical properties. Cell Motil 1983;3:609eC622.

    Evans E, Needham D. Physical properties of surfactant bilayer membranes composed of lipids, cholesterol and polypeptides: thermal transitions, elasticity, cohesion, and colloidal interaction. J Phys Chem 1987;91:4219eC4228.

    Bloom M, Evans E, Mouritsen OG. Physical properties of the fluid bilayer component of cell membranes. Q Rev Biophys 1991;24:293eC397.

    Evans E, Heinrich V, Ludwig F, Rawicz W. Dynamic tension spectroscopy and strength of biomembranes. Biophys J 2003;85:2342eC2350.

    Benz R, Zimmermann U. The resealing process of lipid bilayers after reversible electrical breakdown. Biochim Biophys Acta 1981;640:169eC178.

    Lipowsky R. The conformation of membranes. Nature 1991;349:475eC481.

    McNeil PL, Miyake K, Vogel SS. The endomembrane requirement for cell surface repair. Proc Natl Acad Sci USA 2003;100:4592eC4597.

    Steinhardt RA, Bi G, Alderton JM. Cell membrane resealing by a vesicular mechanism similar to neurotransmitter release. Science 1994;263:390eC393.

    Togo T, Krasieva TB, Steinhardt RA. A decrease in membrane tension precedes successful cell-membrane repair. Mol Biol Cell 2000;11:4339eC4346.

    Togo T, Alderton JM, Bi GQ, Steinhardt RA. The mechanism of facilitated cell membrane resealing. J Cell Sci 1999;112:719eC731.

    Togo T, Steinhardt RA. Nonmuscle myosin IIA and IIB have distinct functions in the exocytosis-dependent process of cell membrane repair. Mol Biol Cell 2004;15:688eC695.

    Terasaki M, Miyake K, McNeil PL. Large plasma membrane disruptions are rapidly resealed by Ca2+-dependent vesicle-vesicle fusion events. J Cell Biol 1997;139:63eC74.

    McNeil PL, Vogel SS, Miyake K, Terasaki M. Patching plasma membrane disruptions with cytoplasmic membrane. J Cell Sci 2000;113:1891eC1902.

    Bi GQ, Alderton JM, Steinhardt RA. Calcium-regulated exocytosis is required for cell membrane resealing. J Cell Biol 1995;131:1747eC1758.

    Bi GQ, Morris RL, Liao G, Alderton JM, Scholey JM, Steinhardt RA. Kinesin- and myosin-driven steps of vesicle recruitment for Ca2+-regulated exocytosis. J Cell Biol 1997;138:999eC1008.

    Reddy A, Caler EV, Andrews NW. Plasma membrane repair is mediated by Ca(2+)-regulated exocytosis of lysosomes. Cell 2001;106:157eC169.

    Jahn R, Lang T, Sudhof TC. Membrane fusion. Cell 2003;112:519eC533.

    McNeil PL. Repairing a torn cell surface: make way, lysosomes to the rescue. J Cell Sci 2002;115:873eC879.

    Andrews NW. Regulated secretion of conventional lysosomes. Trends Cell Biol 2000;10:316eC321.

    Rodriguez A, Webster P, Ortego J, Andrews NW. Lysosomes behave as Ca2+-regulated exocytic vesicles in fibroblasts and epithelial cells. J Cell Biol 1997;137:93eC104.

    Rodriguez A, Martinez I, Chung A, Berlot CH, Andrews NW. cAMP regulates Ca2+-dependent exocytosis of lysosomes and lysosome-mediated cell invasion by trypanosomes. J Biol Chem 1999;274:16754eC16759.

    Jedd G, Chua NH. A new self-assembled peroxisomal vesicle required for efficient resealing of the plasma membrane. Nat Cell Biol 2000;2:226eC231.

    Borgonovo B, Cocucci E, Racchetti G, Podini P, Bachi A, Meldolesi J. Regulated exocytosis: a novel, widely expressed system. Nat Cell Biol 2002;4:955eC962.

    Togo T, Alderton JM, Steinhardt RA. Long-term potentiation of exocytosis and cell membrane repair in fibroblasts. Mol Biol Cell 2003;14:93eC106.

    Rao SK, Huynh C, Proux-Gillardeaux V, Galli T, Andrews NW. Identification of SNAREs involved in synaptotagmin VII-regulated lysosomal exocytosis. J Biol Chem 2004;279:20471eC20479.

    Ninomiya Y, Kishimoto T, Miyashita Y, Kasai H. Ca2+-dependent exocytotic pathways in Chinese hamster ovary fibroblasts revealed by a caged-Ca2+ compound. J Biol Chem 1996;271:17751eC17754.

    Detrait ER, Yoo S, Eddleman CS, Fukuda M, Bittner GD, Fishman HM. Plasmalemmal repair of severed neurites of PC12 cells requires Ca(2+) and synaptotagmin. J Neurosci Res 2000;62:566eC573.

    Chakrabarti S, Kobayashi KS, Flavell RA, Marks CB, Miyake K, Liston DR, Fowler KT, Gorelick FS, Andrews NW. Impaired membrane resealing and autoimmune myositis in synaptotagmin VII-deficient mice. J Cell Biol 2003;162:543eC549.

    Lyerla TA, Rusiniak ME, Borchers M, Jahreis G, Tan J, Ohtake P, Novak EK, Swank RT. Aberrant lung structure, composition, and function in a murine model of Hermansky-Pudlak syndrome. Am J Physiol Lung Cell Mol Physiol 2003;285:L643eCL653.

    Tremblay L, Valenza F, Ribeiro SP, Li J, Slutsky AS. Injurious ventilatory strategies increase cytokines and c-fos m-RNA expression in an isolated rat lung model. J Clin Invest 1997;99:944eC952.

    Dreyfuss D, Ricard JD, Saumon G. On the physiologic and clinical relevance of lung-borne cytokines during ventilator-induced lung injury. Am J Respir Crit Care Med 2003;167:1467eC1471.

    Pugin J, Dunn I, Jolliet P, Tassaux D, Magnenat JL, Nicod LP, Chevrolet JC. Activation of human macrophages by mechanical ventilation in vitro. Am J Physiol 1998;275:L1040eCL1050.

    Vlahakis NE, Schroeder MA, Limper AH, Hubmayr RD. Stretch induces cytokine release by alveolar epithelial cells in vitro. Am J Physiol Lung Cell Mol Physiol 1999;277:L167eCL173.

    Belperio JA, Keane MP, Burdick MD, Londhe V, Xue YY, Li K, Phillips RJ, Strieter RM. Critical role for CXCR2 and CXCR2 ligands during the pathogenesis of ventilator-induced lung injury. J Clin Invest 2002;110:1703eC1716.

    Slutsky AS, Tremblay LN. Multiple system organ failure: is mechanical ventilation a contributing factor Am J Respir Crit Care Med 1998;157:1721eC1725.

    Murphy DB, Cregg N, Tremblay L, Engelberts D, Laffey JG, Slutsky AS, Romaschin A, Kavanagh BP. Adverse ventilatory strategy causes pulmonary-to-systemic translocation of endotoxin. Am J Respir Crit Care Med 2000;162:27eC33.

    Chiumello D, Pristine G, Slutsky AS. Mechanical ventilation affects local and systemic cytokines in an animal model of acute respiratory distress syndrome. Am J Respir Crit Care Med 1999;160:109eC116.

    Nahum A, Hoyt J, Schmitz L, Moody J, Shapiro R, Marini JJ. Effect of mechanical ventilation strategy on dissemination of intratracheally instilled Escherichia coli in dogs. Crit Care Med 1997;25:1733eC1743.

    Verbrugge SJ, Sorm V, van't Veen A, Mouton JW, Gommers D, Lachmann B. Lung overinflation without positive end-expiratory pressure promotes bacteremia after experimental Klebsiella pneumoniae inoculation. Intensive Care Med 1998;24:172eC177.

    Haitsma JJ, Uhlig S, Goggel R, Verbrugge SJ, Lachmann U, Lachmann B. Ventilator-induced lung injury leads to loss of alveolar and systemic compartmentalization of tumor necrosis factor-alpha. Intensive Care Med 2000;26:1515eC1522.

    Imai Y, Parodo J, Kajikawa O, de Perrot M, Fischer S, Edwards V, Cutz E, Liu MY, Keshavjee S, Martin TR, et al. Injurious mechanical ventilation and end-organ epithelial cell apoptosis and organ dysfunction in an experimental model of acute respiratory distress syndrome. JAMA 2003;289:2104eC2112.

    Bouadma L, Schortgen F, Ricard JD, Martet G, Dreyfuss D, Saumon G. Ventilation strategy affects cytokine release after mesenteric ischemia-reperfusion in rats. Crit Care Med 2004;32:1563eC1569.

    Correa-Meyer E, Pesce L, Guerrero C, Sznajder JI. Cyclic stretch activates ERK1/2 via G proteins and EGFR in alveolar epithelial cells. Am J Physiol Lung Cell Mol Physiol 2002;282:L883eCL891.

    Copland IB, Kavanagh BP, Engelberts D, McKerlie C, Belik J, Post M. Early changes in lung gene expression due to high tidal volume. Am J Respir Crit Care Med 2003;168:1051eC1059.

    Bhattacharya S, Sen N, Yiming MT, Patel R, Parthasarathi K, Quadri S, Issekutz AC, Bhattacharya J. High tidal volume ventilation induces proinflammatory signaling in rat lung endothelium. Am J Respir Crit Care Med 2003;28:218eC224.

    Li LF, Ouyang B, Choukroun G, Matyal R, Mascarenhas M, Jafari B, Bonventre JV. Force T, Quinn DA. Stretch-induced IL-8 depends on c-Jun NH2-terminal and nuclear factor-kappaB-inducing kinases. Am J Physiol Lung Cell Mol Physiol 2003;285:L464eCL475.

    Matthay MA, Zimmerman GA, Esmon C, Bhattacharya J, Coller B, Doerschuk CM, Floros J, Gimbrone MA Jr, Hoffman E, Hubmayr RD, et al. Future research directions in acute lung injury: summary of a National Heart, Lung, and Blood Institute working group. Am J Respir Crit Care Med 2003;167:1027eC1035.

    Grembowicz KP, Sprague D, McNeil PL. Temporary disruption of the plasma membrane is required for c-fos expression in response to mechanical stress. Mol Biol Cell 1999;10:1247eC1257.

    Ichimura H, Parthasarathi K, Quadri S, Issekutz AC, Bhattacharya J. Mechano-oxidative coupling by mitochondria induces proinflammatory responses in lung venular capillaries. J Clin Invest 2003;111:691eC699.

    Wang PM, Fujita E, Bhattacharya J. Vascular regulation of type II cell exocytosis. Am J Physiol Lung Cell Mol Physiol 2002;282:L912eCL916.

    Kuebler WM, Ying X, Bhattacharya J. Pressure-induced endothelial Ca(2+) oscillations in lung capillaries. Am J Physiol Lung Cell Mol Physiol 2002;282:L917eCL923.

    Boitano S, Safdar Z, Welsh DG, Bhattacharya J, Koval M. Cell-cell interactions in regulating lung function. Am J Physiol Lung Cell Mol Physiol 2004;287:L455eCL459.

    Martin TR, Nakamura M, Matute-Bello G. The role of apoptosis in acute lung injury. Crit Care Med 2003;31:S184eCS188.

    Del Riccio V, van Tuyl M, Post M. Apoptosis in lung development and neonatal lung injury. Pediatr Res 2004;55:183eC189.

    Nanjundan M, Possmayer F. Pulmonary phosphatidic acid phosphatase and lipid phosphate phosphohydrolase. Am J Physiol Lung Cell Mol Physiol 2003;284:L1eCL23.

    Bardales RH, Xie SS, Schaefer RF, Hsu SM. Apoptosis is a major pathway responsible for the resolution of type II pneumocytes in acute lung injury. Am J Pathol 1996;149:845eC852.

    Maniotis AJ, Chen CS, Ingber DE. Demonstration of mechanical connections between integrins, cytoskeletal filaments, and nucleoplasm that stabilize nuclear structure. Proc Natl Acad Sci USA 1997;94:849eC854.

    Sotoudeh M, Li YS, Yajima N, Chang CC, Tsou TC, Wang Y, Usami S, Ratcliffe A, Chien S, Shyy JY. Induction of apoptosis in vascular smooth muscle cells by mechanical stretch. Am J Physiol Heart Circ Physiol 2002;282:H1709eCH1716.

    Persoon-Rothert M, van der Wees KG, van der Laarse A. Mechanical overload-induced apoptosis: a study in cultured neonatal ventricular myocytes and fibroblasts. Mol Cell Biochem 2002;241:115eC124.

    Galvin DJ, Watson RW, Gillespie JI, Brady H, Fitzpatrick JM. Mechanical stretch regulates cell survival in human bladder smooth muscle cells in vitro. Am J Physiol Renal Physiol 2002;283:F1192eCF1199.

    Edwards YS, Sutherland LM, Power JH, Nicholas TE, Murray AW. Cyclic stretch induces both apoptosis and secretion in rat alveolar type II cells. FEBS Lett 1999;448:127eC130.

    Edwards YS, Sutherland LM, Murray AW. NO protects alveolar type II cells from stretch-induced apoptosis. A novel role for macrophages in the lung. Am J Physiol Lung Cell Mol Physiol 2000;279:L1236eCL1242.

    Upadhyay D, Correa-Meyer E, Sznajder JI, Kamp DW. FGF-10 prevents mechanical stretch-induced alveolar epithelial cell DNA damage via MAPK activation. Am J Physiol Lung Cell Mol Physiol 2003;284:L350eCL359.

    Putensen C, Hering R, Wrigge H. Controlled versus assisted mechanical ventilation. Curr Opin Crit Care 2002;8:51eC57.

    Mascheroni D, Kolobow T, Fumagalli R, Moretti MP, Chen V, Buckhold D. Acute respiratory failure following pharmacologically induced hyperventilation: an experimental animal study. Intensive Care Med 1988;15:8eC14.

    Veldhuizen RA, Welk B, Harbottle R, Hearn S, Nag K, Petersen N, Possmayer F. Mechanical ventilation of isolated rat lungs changes the structure and biophysical properties of surfactant. J Appl Physiol 2002;92:1169eC1175.

    Veldhuizen RA, Yao LJ, Lewis JF. An examination of the different variables affecting surfactant aggregate conversion in vitro. Exp Lung Res 1999;25:127eC141.

    Ito Y, Veldhuizen RA, Yao LJ, McCaig LA, Bartlett AJ, Lewis JF. Ventilation strategies affect surfactant aggregate conversion in acute lung injury. Am J Respir Crit Care Med 1997;155:493eC499.

    Parker JC, Ivey CL. Isoproterenol attenuates high vascular pressure-induced permeability increases in isolated rat lungs. J Appl Physiol 1997;83:1962eC1967.

    Tremblay LN, Slutsky AS. Ventilator-induced injury: from barotrauma to biotrauma. Proc Assoc Am Physicians 1998;110:482eC488.

作者: Nicholas E. Vlahakis and Rolf D. Hubmayr 2007-5-14
医学百科App—中西医基础知识学习工具
  • 相关内容
  • 近期更新
  • 热文榜
  • 医学百科App—健康测试工具