Skip to main content
NIHPA Author Manuscripts logoLink to NIHPA Author Manuscripts
. Author manuscript; available in PMC: 2011 Jan 5.
Published in final edited form as: J Biomech. 2009 Oct 4;43(1):99. doi: 10.1016/j.jbiomech.2009.09.015

Recent Advances and New Opportunities in Lung Mechanobiology

Daniel J Tschumperlin 1, Francis Boudreault 1, Fei Liu 1
PMCID: PMC2813412  NIHMSID: NIHMS147094  PMID: 19804885

Abstract

Lung function is inextricably linked to mechanics. On short timescales every breath generates dynamic cycles of cell and matrix stretch, along with convection of fluids in the airways and vasculature. Perturbations such airway smooth muscle shortening or surfactant dysfunction rapidly alter respiratory mechanics, with profound influence on lung function. On longer timescales, lung development, maturation, and remodeling all strongly depend on cues from the mechanical environment. Thus mechanics has long played a central role in our developing understanding of lung biology and respiratory physiology. This concise review focuses on progress over the past five years in elucidating the molecular origins of lung mechanical behavior, and the cellular signaling events triggered by mechanical perturbations that contribute to lung development, homeostasis, and injury. Special emphasis is placed on the tools and approaches opening new avenues for investigation of lung behavior at integrative cellular and molecular scales. We conclude with a brief summary of selected opportunities and challenges that lie ahead for the lung mechanobiology research community.

Keywords: mechanotransduction, extracellular matrix, respiratory, stretch

The evolution of lung mechanobiology

The core function of the lung, to facilitate the exchange of gases between the circulation and the external environment, is inextricably linked with mechanics. Respiratory muscles generate a transpulmonary pressure gradient, prompting gas to flow through the branched structure of the airways to alveoli whose stability depends on a fine balance of tissue and surface forces (Fredberg and Kamm 2006), while blood from the heart circulates through a dense network of capillaries to exchange CO2 and O2 across the delicate alveolar-capillary walls (Maina and West 2005). Understanding the physical origins of these functions, and their failure in various disease states, has been central to the study of respiratory physiology and medicine since the inception of these fields. Seminal historical developments, such as the characterization of pulmonary surfactant function (Obladen 2005), helped to reveal the central role of mechanics in lung function. Current mechanobiological study of the lung thus builds on a long history and a rich foundation. This review focuses on recent progress over the preceding five years, emphasizing new tools and approaches driving progress in the field, and new insights into molecular, cellular and microstructural aspects of the biology-mechanics interface helping to inform our evolving understanding of lung function in health and disease.

Imaging tools open new doors

Mechanobiological study of the lung requires a detailed understanding of the mechanical state of the tissue under physiological conditions. The delicate and complex microstructure of the lung, encased within the thoracic cavity, has posed a long-standing challenge in terms of access for imaging and measurement. Traditionally, measurements of lung mechanics and attribution of mechanical contributions to various anatomical or tissue components have been inferred from pressure, volume, and flow relationships obtained at the entrance to the airways. While great strides have been made in extracting information from these measurements, new advances in methods to visualize the microstructure and dynamics of living lung tissue are opening exciting new opportunities.

At the macro-scale, application of non-invasive imaging modalities are providing new information about regional tissue deformations and may make possible local measurements of intact lung mechanical properties. For instance, microfocal x-ray imaging of airways is allowing unprecedented measurements of airway dimensions in intact lungs (Fig. 1A), providing new insight into regional, axial and circumferential variations in airway strains that occur with changing lung volume (Sera et al. 2004; Sinclair et al. 2007). Magnetic resonance elastography (MRE) is a technique that is well-established as a non-invasive means to sample tissue mechanics in soft organs such as the liver. While application of MRE to the lung is complicated by its air-filled structure, preliminary work using porcine lungs inflated with hyperpolarized 3He validates the feasibility of applying this methodology to measure tissue mechanical properties within intact lungs (McGee et al. 2008).

Figure 1.

Figure 1

(A) Micro-CT images of rat lung airways at inflation volumes of (a) functional residual capacity (FRC) and (b) total lung capacity (TLC). Same-direction arrows indicate the same airways. Scale bar, 500 μm. Adapted with permission from (Sera et al. 2004). (B) The same rat lung alveolus imaged with intravital microscopy at transpulmonary inflation pressures of 5 cmH2O (green pseudocolor) and 20 cmH2O (red pseudocolor). Numbers in baseline image label two perimeter segments. An overlay of the images demonstrates inflation-induced alveolar expansion, which increased total alveolar perimeter length, L and alveolar diameter, D by 13 and 15%, respectively. Adapted with permission from (Perlman and Bhattacharya 2007).

At the micro-scale, recent advances have brought the power of microscopy to the visualization of internal lung microstructure. While limited to the peripheral subpleural region of lung tissue, intravital microscopy has opened new opportunities to visualize events within the alveoli and microvasculature of the intact lung (Kuebler et al. 2007). Exciting opportunities now exist to couple such imaging tools with fluorescence indicators of cellular signaling events allowing the study of mechanotransduction in situ (Kuebler et al. 2007; Sabouri-Ghomi et al. 2008). Moreover, intravital imaging approaches are leading the way towards an increasingly sophisticated understanding of local mechanical properties and deformations in the intact lung (Popp et al. 2006; Perlman and Bhattacharya 2007).

One example of the way intravital microscopy has already enriched our understanding of local intra-alveolar deformations comes from confocal imaging of subpleural alveoli in isolated rat lungs (Perlman and Bhattacharya 2007). This study demonstrated striking heterogeneity in the behavior of alveolar wall segments (Fig. 1B), with greater distention observed in wall segments associated with type I epithelial cells than in those segments associated with type II epithelial cells (Perlman and Bhattacharya 2007). Type II cells are more vulnerable to cell injury and or death when stretched than are type I cells (Tschumperlin and Margulies 1998). Together these findings suggest that the alveolus is designed to protect type II cells from large distention, either through geometric or compositional effects on the underlying matrix, or through changes in the stiffness of the individual cell types themselves. Support for the latter comes from application of atomic force microscopy (AFM) to isolated lung epithelial cells, where it was observed that the cytoplasm of type II cells is stiffer (~2-fold median difference) than that of type I cells (Azeloglu et al. 2008).

AFM, while more invasive than the methods described above, provides a unique perspective because it can characterize cell and tissue mechanical properties with micron-scale spatial resolution. Recently, AFM was used to map the local elasticity of migrating epithelial cells in a model of wound healing (Wagh et al. 2008), demonstrating the great potential for AFM to probe cellular and subcellular mechanics in isolated cells (Gavara et al. 2008; Kang et al. 2008). But AFM has also been used to characterize the local mechanical properties of intact soft tissues (Engler et al. 2004; Berry et al. 2006; Engler et al. 2007). We have employed AFM microindentation to spatially map the stiffness of lung tissue slices harvested from mice with bleomycin-induced fibrosis (Fig. 2A, unpublished data). Prior measurements of lung tissue at the macro-scale indicated an approximate doubling or tripling of tissue stiffness with fibrosis (Dolhnikoff et al. 1999; Ebihara et al. 2000). In contrast, AFM microindentation reveals that tissue stiffening is highly localized, with some regions up to ~30-fold stiffer than the median observed in normal lung tissue. Given the profound effects matrix stiffness can exert on fibroblasts (Wang et al. 2000; Goffin et al. 2006; Li et al. 2007; Wipff et al. 2007), it's likely that lung remodeling and changes in matrix rigidity play a prominent role in various pathologies of the lung, including asthmatic airway remodeling, fibrosis, emphysema, and pulmonary hypertension. So far the role of matrix stiffness in the initiation and amplification of these lung diseases remains largely unexplored.

Figure 2.

Figure 2

(A) Representative stiffness maps of normal and fibrotic (14 days post bleomycin treatment) mouse lung parenchyma. Colorbar indicates shear modulus. AFM force-indentation profiles were acquired in a 16×16 sample grid separated by 5 μm spatially covering 80×80 μm area. Shear modulus at each point on the grid was calculated from fitting force-indentation data using a Hertz sphere model and resulting shear modulus data were plotted in a contour map (unpublished data). (B) Simulation of the progression of pulmonary fibrosis (a) and emphysema (b) based on percolation of sequential alveolar wall stiffening or rupture. (a) The curve shows the bulk modulus of the elastic network versus the fraction of springs randomly stiffened by a factor of 100. If all the spring constants were uniformly stiffened in a gradual manner from the baseline value of 1 to 100, the modulus would follow the dashed diagonal line. Top: Network configurations obtained when 0, 50, and 67% of the springs have been stiffened. (b) The curve shows the bulk modulus of the elastic network versus the fraction of springs cut on the basis of the amount of tension they carry. Top: Network configurations obtained at three points along this process. The stresses in the individual springs are indicated by color coding, with yellow indicating high stress and decreasing stress corresponding to progressively darker shades of blue. Adapted with permission from (Bates et al 2007).

Micro-scale and multi-scale mechanics

A major goal of lung mechanobiology is to understand how the mechanical behavior of the lung emerges from its molecular and cellular constituents (Suki and Bates 2008). Cell-matrix model systems have been used to identify the stimuli that promote cell-mediated matrix remodeling, and concomitant changes in matrix organization and mechanical behavior (Leung et al. 2007; Raub et al. 2007; Raub et al. 2008). Analysis of second harmonic generation from multi-photon imaging of cell-matrix constructs (Raub et al. 2007; Raub et al. 2008) is being used to explore linkages between structural and mechanical properties of collagen matrices. Mechanical testing of lung tissue strips, coupled to selective perturbation of matrix constituents, is advancing our understanding of how collagen, elastin and other matrix proteins contribute to tissue stability and deformability (Kononov et al. 2001; Cavalcante et al. 2005; Jesudason et al. 2007). Similarly, lungs from mice genetically deficient for the proteoglycan decorin support a prominent role of this matrix protein in lung mechanics (Fust et al. 2005). These studies are challenging the long held assumption that lung mechanics can be simply partitioned into contributions from elastin at low volume, and collagen at high volume. Based on compelling evidence that alveolar walls can fail under loading, particularly when the matrix is remodeled (Kononov et al. 2001; Ito et al. 2004; Ito et al. 2005; Ito et al. 2006; Ritter et al. 2009), Suki and colleagues have put forward a compelling hypothesis (Fig. 2B) to explain progressive emphysema based on percolation of sequential alveolar wall rupture (Suki et al. 2005; Bates et al. 2007; Suki and Bates 2008). Together these approaches are building a multi-level hierarchical understanding of lung tissue mechanics, moving toward the ultimate goal of predictive power to understand how molecular perturbations alter lung micromechanics.

In parallel with advances in solid mechanics, ongoing investigations of biofluid mechanics in the pulmonary system (Bertram and Gaver 2005) are providing an improved understanding of lung injury mechanisms associated with closure and opening of fragile tissue structures (Kay et al. 2004; Yalcin et al. 2007), and generating multi-scale models for investigation of fluid dynamics in airways and lung vasculature (Tawhai and Burrowes 2008). At the interface of solid and fluid surfaces, recent experimental work has demonstrated that the airway-lining layer of mucus, which is optimized for lung defense and clearance, is responsive to airway shear stresses (Tarran et al. 2005; Tarran et al. 2006), and that pathological changes in mucus viscoelasticity can be targeted as a new therapeutic modality for patients with cystic fibrosis (Donaldson et al. 2006; Tarran et al. 2007).

Flow resistance through the airways is highly responsive to the dimensions of airways, which can be dynamically regulated by airway smooth muscle contraction. Because airway narrowing compromises efficient gas transport, the dysregulation of airway dimensions represents a critical pathogenetic mechanism in both asthma and chronic obstructive pulmonary disease (COPD). Over the past decade the demonstration that airway smooth muscle tone is dynamically equilibrated, and thus sensitive to lung volume history, represents a major advance in lung mechanobiology (An et al. 2007). However, fundamental questions remain regarding mechanisms of airways hyperresponsiveness, and whether the locus of hyperresponsiveness is the smooth muscle itself, the remodeled airway, matrix or cellular constituents, or some combination (McParland et al. 2003; Wagers et al. 2004; Bates and Lauzon 2007; James and Wenzel 2007; Wagers et al. 2007).

Stretching the lung

The effects of stretch, as the central physical change required for lung inflation, continues to dominate lung mechanobiological investigation at organ and cellular scales. At the whole organ scale the effects of stretch are manifested in three major settings: lung development, compensatory growth, and injury.

During lung development in utero the epithelium is secretory, in contrast to its absorptive role in mature lungs. Lumenal secretions across the epithelium flow through the developing airways and exit through the larynx and nasopharynx where the partial occlusion of the vocal cords acts as a one way valve that generates back pressure to partially inflate the growing lungs. This tonic inflation is critical to lung development, as failure to inflate retards lung growth and maturation (Tschumperlin and Drazen 2006). Tracheal occlusion increases lumenal expansion, and accelerates branching and cellular maturation (Unbekandt et al. 2008). Recent experiments using organ cultures of primitive lung buds and perturbations targeting actin-myosin contractility demonstrate a critical role for local force generation in branching morphogenesis (Moore et al. 2002; Moore et al. 2005). While the organ-level response to tracheal occlusion is well documented, the dissection of the cellular events that follow tracheal occlusion is only recently begun with the demonstration of rapid increases in proliferation of both epithelial and mesenchymal cells (Seaborn et al. 2008). Further efforts directed at elucidating the molecular mechanisms coupling tracheal occlusion to accelerated airway branching point to the FGF10-FGFR2b signaling axis (Unbekandt et al. 2008). Similar to effects on airway branching, lung distention has also recently been shown to play a pivotal role in coordinating angiogenesis in the developing lung (Cloutier et al. 2008).

Based on the essential role for mechanical forces in lung development, it is not surprising that mechanical forces also play a leading role in the compensatory growth of the lung following surgical resection (2004; Hsia 2004). Thus harnessing the mechanical environment therapeutically could represent a powerful stimulus to treat hypoplastic lungs (Butter et al. 2005) or regenerate functional lung units lost to aging and disease (Fehrenbach et al. 2008). Unfortunately, the robust compensatory growth seen in rodents and young animals is much weaker in adults (2004), and attempts to augment compensatory growth have met with mostly frustrating results (Dane et al. 2004; Ravikumar et al. 2007). The recent description of resident stem cells in a number of lung niches (Rawlins and Hogan 2006; Stripp and Shapiro 2006; Kim 2007), and the possible connection between mechanical stimulation and stem cell activation (Nolen-Walston et al. 2008) offers some hope that mechanobiological approaches may someday help to unlock the lung's inherent regenerative capacity.

On the flip-side of regeneration, a vast body of literature demonstrates potentially pathological responses to lung stretch (Oeckler and Hubmayr 2007), culminating in clinical trials demonstrating protective effects of reducing tidal volume during mechanical ventilation (2000; Meade et al. 2008). Recently, experimental studies have explored how even moderate volume ventilation, when superimposed on existing pulmonary conditions, can result in deleterious effects on lung function (Altemeier et al. 2004; Altemeier et al. 2005; Bregeon et al. 2005; Tsuchida et al. 2005; Dhanireddy et al. 2006; Levine et al. 2006; O'Mahony et al. 2006; Tsuchida et al. 2006).

Intense research efforts continue into the molecular mechanisms that couple ventilatory stresses and strains into adverse physiological outcomes. Recently, TRPV4 (Hamanaka et al. 2007) PI3K, Akt and Src (Miyahara et al. 2007) have been implicated in capillary leakage, while hyaluronan fragmentation was implicated in IL-8 upregulation (Mascarenhas et al. 2004), and cellular stress failure demonstrated in lung over-distention injury (Vlahakis and Hubmayr 2005). Neutrophils play a prominent role in acute lung injury and their activation by injurious mechanical ventilation strategies continues to be an area of active investigation, with stretch induced stiffening of neutrophils and signaling through c-Jun N-terminal kinase (Choudhury et al. 2004; Li et al. 2004) both implicated in neutrophil margination and activation in the lung microcirculation. The observation that ventilation may result in profoundly different outcomes in neo-natal versus adult populations has led to the development of comparative animal ventilation models which should help to uncover the developmental basis for differential responses to lung distention (Copland et al. 2004; Kornecki et al. 2005). While hypothesis-driven approaches provide mechanistic dissection of ventilation-induced biochemical signaling, discovery based approaches have also been implemented to survey the landscape of transcriptional events initiated by stretch. Microarray analyses highlight genomic scale events, transcriptional programs and candidate genes evoked by the ventilation in various model systems (Grigoryev et al. 2004; Altemeier et al. 2005; Ma et al. 2005; Dolinay et al. 2006; Gharib et al. 2006; Simon et al. 2006; Wurfel 2007; dos Santos et al. 2008). Ultimately, the challenge will be to reverse engineer the mechanobiological pathways underlying these transcriptional programs, and develop strategies based on this knowledge to avoid or attenuate adverse responses to mechanical ventilation.

Stretching lung cells

In parallel with animal models of ventilation, cell culture approaches are being widely exploited to accelerate discovery and characterization of mechanical signaling events in lung cells. While limited in their capacity to recapitulate in vivo biology and the complex intercellular interactions that are present in the lung, these models provide enhanced control of the mechanical and biochemical environment, and offer ready opportunities to interrogate specific cellular and molecular events modulated by mechanical stimuli.

Many of the cellular mechanobiology studies relevant to the lung focus on the epithelium, which covers the vast surface area of the lung, and thus represents a potent source for mechanical regulation. Not surprisingly, studies of distal lung epithelium largely parallel whole lung studies with interrelated emphases on development and differentiation, growth responses, and injury mechanisms. Studies of fetal lung cells have identified a vast array of signaling responses to stretch (Copland and Post 2007), as well as secretion of paracrine mediators such as prostanoids (Copland et al. 2006) and serotonin (Pan et al. 2006). Confirming the stimulatory effects of stretch on accelerated lung maturation and cellular differentiation, cell culture studies demonstrate the potent effects of stretch on fetal type II epithelial differentiation, with contributions from EGFR (Sanchez-Esteban et al. 2004), integrins β1, α6, and α3 (Sanchez-Esteban et al. 2006), and cAMP-PKA signaling (Wang et al. 2006). Global microarray analysis of fetal type II epithelium exposed to stretch demonstrated enhanced expression of a variety of genes, including the amiloride-sensitive epithelial sodium channel gene (Scnn1a), suggesting a role for stretch in preparing the developing epithelium for the transition from net secretion to absorption at birth (Wang et al. 2006). Together these studies provide testable new hypotheses for critical regulatory nodes in mechanical regulation of lung development.

In cells representative of those in the mature lung, stretch has been shown to exert potent effects on growth, structure, homeostasis and differentiated function. Cyclic stretch drives the production of reactive oxygen species (ROS) in distal lung epithelial cells (Chapman et al. 2005) necessary for stretch-induced cellular proliferation (Chess et al. 2005). The growth regulation of epithelial cells by stretch also requires contributions from Src, FAK, and ERK, emphasizing the complex interplay of regulatory signals governing cellular proliferation in response to mechanical stimulation (Chaturvedi et al. 2007). A novel dystroglycan-dependent mechanotransduction response has been observed in stretched alveolar cells (Jones et al. 2005), and contributes to activation of AMP kinase and ERK (Jones et al. 2005; Budinger et al. 2008). Stretch also leads to remodeling of the epithelial intermediate filament network (Felder et al. 2008) and actin cytoskeleton (Papaiahgari et al. 2007), and calcium-dependent fusion of lamellar bodies with the cell membrane (Frick et al. 2004). Lamellar bodies are the densely packed structures used by cells to store and deliver the phospholipid components of pulmonary surfactant. Interestingly, mixed cultures of type I and type II cells respond to stretch with greater phospholipid secretion than type II cells alone, implicating type I cells as primary mechanosensors which stimulate secretion in neighboring type II cells though extracellular ATP signaling (Patel et al. 2005).

As in the intact lung, not all effects of stretch are beneficial or benign for distal lung epithelial cells. Cyclic stretch causes an acidification response that promotes bacterial growth (Pugin et al. 2008), impairs the barrier function of epithelial monolayers (Cavanaugh et al. 2006), and leads to cell death and release of cytokines that play important roles in ventilator-induced lung injury (Hammerschmidt et al. 2004; Hammerschmidt et al. 2005; Hammerschmidt et al. 2007). Cell culture models also recapitulate the observation that moderate stretch alone is less provocative of cellular responses than is stretch in the presence of an infectious agent (LPS) or inflammatory cytokine (TNF-alpha) (dos Santos et al. 2004). Intriguingly, there is one example of using cell culture models to show that the epithelium can be treated to enhance its resistance to the negative effects of stretch, with compelling evidence that IL-10 has protective effects worthy of follow-on study in animal models (Lee et al. 2008). Similarly, experimental and computational approaches have been employed to study stretch-induced increase in Na-K-ATPase pumping activity (Fisher and Margulies 2007), which could be a target to improve fluid clearance from airspaces in acute lung injury.

For the proximal airways, air-liquid interface cultures of primary bronchial epithelium faithfully recapitulate the differentiated character of this tissue, and have been used to elucidate both ATP- and EGFR-dependent modes of epithelial mechanotransduction (Tschumperlin et al. 2004; Kojic et al. 2006; Button et al. 2007). The EGFR-dependent response is evoked by compressive stress shrinking the lateral intercellular space (Fig. 3A-B), and has been further linked to positive feedback signal amplification and activation of the plasminogen family of enzymes that exert broad control over matrix remodeling pathways (Chu et al. 2005; Chu et al. 2006). Mechanotransduction of chronic intermittent compressive stress (Fig. 3C) also enhances expression of mucin protein (Park and Tschumperlin 2009), suggesting a causal link between mechanical perturbations in the airways and excessive mucus production that contributes to airway obstruction. Further effort has led to the development of co-culture models which incorporate differentiated epithelium adjacent to cell-populated collagen matrices, allowing for cell-cell communication in configurations approximating the native airway wall (Choe et al. 2006). These systems demonstrate an integrated tissue remodeling response to mechanical stress that resembles many of the features of asthmatic airway remodeling (Swartz et al. 2001).

Figure 3.

Figure 3

(A) Schematic of bronchial epithelial cells cultured at air-liquid interface on a microporous substrate. The lateral cellular surfaces express pro-ligands of the EGF family and their cognate EGFR receptors, forming a local autocrine circuit. (B) Compressive stress (apical to basal transcellular pressure gradient) shrinks the lateral intercellular space between neighboring bronchial epithelial cells, visualized by two-photon imaging of extracellular fluorescent dextran. Sequential images at baseline (0 seconds), 60 and 600 seconds after initiation of continuous compressive stress illustrate the gradual decline in intercellular gap distance. (C) Chronic intermittent exposure to compressive stress daily for 14 days enhances expression of a mucus secretory phenotype, visualized by immunofluorescent staining of MUC5AC (green). Nuclear counterstain is shown in red.

Opportunities and Challenges

The new tools and methods being deployed for mechanobiology investigation in the lung are accelerating the pace of discovery, and creating opportunities for breakthroughs. While the challenges are many, the field appears poised for tackle longstanding questions. Furthermore, the chance to translate basic understanding into clinical practice and therapies is real and closer than ever before. Below we present several thematic areas that, from our perspective, appear ripe for discovery, synthesis, and innovation.

Throughout the text the diverse and profound effects of lung distention in vivo, and stretch in vitro have been emphasized. Unresolved to this day is the paradoxical observation that stretch is necessary for lung growth and regeneration, but also capable of causing or exacerbating injury in the context of mechanical ventilation. While this paradox might be dismissed as an issue of degree, with small distention promoting growth, and large distention causing injury, several lines of evidence indicate that the situation is more complex. Moderate to large lung distention, in the absence of preexisting injury or infection, can be quite well tolerated (Hsia 2004). In contrast, even modest degrees of lung distention can cause or amplify injury when applied to lungs with local or systemic inflammation (Altemeier et al. 2004; Altemeier et al. 2005). The interaction between the immune system, the existing soluble milieu, and the cellular responses to stretch would appear to constitute a highly complex, but extraordinarily important intersection. If critical mediators or pathways could be identified to shift cellular stretch responses from those associated with injury and inflammatory signaling toward those responses associated with growth and regeneration, the implications for both mechanical ventilation and lung regeneration would be profound.

Similarly, nascent efforts to engineer lung tissue (Hoganson et al. 2008; Nichols and Cortiella 2008; Zani et al. 2008), or develop scaffolds for lung tissue regeneration would appear natural beneficiaries of enhanced understanding of the key pathways that promote growth and regeneration responses to lung distention. Here the investigation of lung development, and the differing responses of neonatal and mature lungs to mechanical perturbations could be employed to uncover critical pathways selectively activated in lung growth (Copland et al. 2004; Kornecki et al. 2005). As knowledge of progenitor cell function in the lung advances, the effects of the mechanical environment on these cells will also need to be elucidated. For instance, it will be interesting to define whether the mechanical environment is one of the characteristics that defines niches for progenitor cells, or plays any role in commitment of progenitor cells to various lung-specific lineages.

While stretch has been, and will continue to be a focal point for mechanobiological investigation in the lung, the role of stiffness appears poised for increased recognition. The lung parenchyma is a highly compliant tissue, and prominent diseases such as pulmonary fibrosis, emphysema, and asthmatic airway remodeling alter lung tissue compliance. The role that stiffness plays in initiating, amplifying, or prolonging these disease processes represents fertile but as yet unexplored territory. Moreover, the already discussed in vitro effects of stretch may need to be re-evaluated with regards to stiffness – it is not known if stretching cells on soft, physiologic stiffness substrates, as opposed to the much stiffer elastic substrates currently employed, will alter cell responses to stretch. In the lung, the amount of tissue stretch is linked to the local and global tissue stiffness, thus as disease processes remodel the lung the tonic and cyclic levels of local distention will simultaneously change. The development of in vitro models capable of simultaneously modulating stretch and stiffness will be needed to explore the intersection of these key aspects of the mechanical environment.

Finally, the search for mechanisms of mechanotransduction in the lung, which has been quite successful over the past several years, has led to an unexpected question: what cellular processes and signaling pathways are unaffected by stretch (and perhaps stiffness)? While clearly there are cell-specific and stimulus specific mechanotransduction responses, it is also increasingly clear that mechanical events can have wide-ranging and potent effects across a variety of pathways and cell types (Trepat et al. 2007). We suggest that the field could benefit from applying integrative and systems type approaches (Janes et al. 2005; Miller-Jensen et al. 2007; Simpson et al. 2008) to dissect critical mechanoregulated events in the lung and lung cells. Such an approach appears especially appealing in light of the diversity of responses known to occur downstream from lung distention, and the divergent physiological responses that can emerge when similar stimuli are applied in different contexts. The application of genomics, proteomics, and systems biology tools and approaches will be increasingly necessary as we move toward a more integrated understanding of mechanobiolgical processes in lung health and disease.

Acknowledgments

The authors acknowledge grant support from the NIH (HL082856 and HL088028) and the Scleroderma Foundation.

Footnotes

Publisher's Disclaimer: This is a PDF file of an unedited manuscript that has been accepted for publication. As a service to our customers we are providing this early version of the manuscript. The manuscript will undergo copyediting, typesetting, and review of the resulting proof before it is published in its final citable form. Please note that during the production process errors may be discovered which could affect the content, and all legal disclaimers that apply to the journal pertain.

References

  1. Ventilation with lower tidal volumes as compared with traditional tidal volumes for acute lung injury and the acute respiratory distress syndrome. The Acute Respiratory Distress Syndrome Network. N Engl J Med. 2000;342:1301–8. doi: 10.1056/NEJM200005043421801. [DOI] [PubMed] [Google Scholar]
  2. Mechanisms and limits of induced postnatal lung growth. Am J Respir Crit Care Med. 2004;170:319–43. doi: 10.1164/rccm.200209-1062ST. [DOI] [PubMed] [Google Scholar]
  3. Altemeier WA, Matute-Bello G, Frevert CW, Kawata Y, Kajikawa O, Martin TR, Glenny RW. Mechanical ventilation with moderate tidal volumes synergistically increases lung cytokine response to systemic endotoxin. Am J Physiol Lung Cell Mol Physiol. 2004;287:L533–42. doi: 10.1152/ajplung.00004.2004. [DOI] [PubMed] [Google Scholar]
  4. Altemeier WA, Matute-Bello G, Gharib SA, Glenny RW, Martin TR, Liles WC. Modulation of lipopolysaccharide-induced gene transcription and promotion of lung injury by mechanical ventilation. J Immunol. 2005;175:3369–76. doi: 10.4049/jimmunol.175.5.3369. [DOI] [PubMed] [Google Scholar]
  5. An SS, Bai TR, Bates JH, Black JL, Brown RH, Brusasco V, Chitano P, Deng L, Dowell M, Eidelman DH, Fabry B, Fairbank NJ, Ford LE, Fredberg JJ, Gerthoffer WT, Gilbert SH, Gosens R, Gunst SJ, Halayko AJ, Ingram RH, Irvin CG, James AL, Janssen LJ, King GG, Knight DA, Lauzon AM, Lakser OJ, Ludwig MS, Lutchen KR, Maksym GN, Martin JG, Mauad T, McParland BE, Mijailovich SM, Mitchell HW, Mitchell RW, Mitzner W, Murphy TM, Pare PD, Pellegrino R, Sanderson MJ, Schellenberg RR, Seow CY, Silveira PS, Smith PG, Solway J, Stephens NL, Sterk PJ, Stewart AG, Tang DD, Tepper RS, Tran T, Wang L. Airway smooth muscle dynamics: a common pathway of airway obstruction in asthma. Eur Respir J. 2007;29:834–60. doi: 10.1183/09031936.00112606. [DOI] [PMC free article] [PubMed] [Google Scholar]
  6. Azeloglu EU, Bhattacharya J, Costa KD. Atomic force microscope elastography reveals phenotypic differences in alveolar cell stiffness. J Appl Physiol. 2008;105:652–61. doi: 10.1152/japplphysiol.00958.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  7. Bates JH, Davis GS, Majumdar A, Butnor KJ, Suki B. Linking parenchymal disease progression to changes in lung mechanical function by percolation. Am J Respir Crit Care Med. 2007;176:617–23. doi: 10.1164/rccm.200611-1739OC. [DOI] [PMC free article] [PubMed] [Google Scholar]
  8. Bates JH, Lauzon AM. Parenchymal tethering, airway wall stiffness, and the dynamics of bronchoconstriction. J Appl Physiol. 2007;102:1912–20. doi: 10.1152/japplphysiol.00980.2006. [DOI] [PubMed] [Google Scholar]
  9. Berry MF, Engler AJ, Woo YJ, Pirolli TJ, Bish LT, Jayasankar V, Morine KJ, Gardner TJ, Discher DE, Sweeney HL. Mesenchymal stem cell injection after myocardial infarction improves myocardial compliance. Am J Physiol Heart Circ Physiol. 2006;290:H2196–203. doi: 10.1152/ajpheart.01017.2005. [DOI] [PubMed] [Google Scholar]
  10. Bertram CD, Gaver DP., 3rd Bio-fluid mechanics of the pulmonary system. Ann Biomed Eng. 2005;33:1681–8. doi: 10.1007/s10439-005-8758-0. [DOI] [PubMed] [Google Scholar]
  11. Bregeon F, Delpierre S, Chetaille B, Kajikawa O, Martin TR, Autillo-Touati A, Jammes Y, Pugin J. Mechanical ventilation affects lung function and cytokine production in an experimental model of endotoxemia. Anesthesiology. 2005;102:331–9. doi: 10.1097/00000542-200502000-00015. [DOI] [PubMed] [Google Scholar]
  12. Budinger GR, Urich D, DeBiase PJ, Chiarella SE, Burgess ZO, Baker CM, Soberanes S, Mutlu GM, Jones JC. Stretch-induced activation of AMP kinase in the lung requires dystroglycan. Am J Respir Cell Mol Biol. 2008;39:666–72. doi: 10.1165/rcmb.2007-0432OC. [DOI] [PMC free article] [PubMed] [Google Scholar]
  13. Butter A, Piedboeuf B, Flageole H, Meehan B, Laberge JM. Postnatal pulmonary distension for the treatment of pulmonary hypoplasia: pilot study in the neonatal piglet model. J Pediatr Surg. 2005;40:826–31. doi: 10.1016/j.jpedsurg.2005.01.058. [DOI] [PubMed] [Google Scholar]
  14. Button B, Picher M, Boucher RC. Differential effects of cyclic and constant stress on ATP release and mucociliary transport by human airway epithelia. J Physiol. 2007;580:577–92. doi: 10.1113/jphysiol.2006.126086. [DOI] [PMC free article] [PubMed] [Google Scholar]
  15. Cavalcante FS, Ito S, Brewer K, Sakai H, Alencar AM, Almeida MP, Andrade JS, Jr., Majumdar A, Ingenito EP, Suki B. Mechanical interactions between collagen and proteoglycans: implications for the stability of lung tissue. J Appl Physiol. 2005;98:672–9. doi: 10.1152/japplphysiol.00619.2004. [DOI] [PubMed] [Google Scholar]
  16. Cavanaugh KJ, Cohen TS, Margulies SS. Stretch increases alveolar epithelial permeability to uncharged micromolecules. Am J Physiol Cell Physiol. 2006;290:C1179–88. doi: 10.1152/ajpcell.00355.2004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  17. Chapman KE, Sinclair SE, Zhuang D, Hassid A, Desai LP, Waters CM. Cyclic mechanical strain increases reactive oxygen species production in pulmonary epithelial cells. Am J Physiol Lung Cell Mol Physiol. 2005;289:L834–41. doi: 10.1152/ajplung.00069.2005. [DOI] [PubMed] [Google Scholar]
  18. Chaturvedi LS, Marsh HM, Basson MD. Src and focal adhesion kinase mediate mechanical strain-induced proliferation and ERK1/2 phosphorylation in human H441 pulmonary epithelial cells. Am J Physiol Cell Physiol. 2007;292:C1701–13. doi: 10.1152/ajpcell.00529.2006. [DOI] [PubMed] [Google Scholar]
  19. Chess PR, O'Reilly MA, Sachs F, Finkelstein JN. Reactive oxidant and p42/44 MAP kinase signaling is necessary for mechanical strain-induced proliferation in pulmonary epithelial cells. J Appl Physiol. 2005;99:1226–32. doi: 10.1152/japplphysiol.01105.2004. [DOI] [PubMed] [Google Scholar]
  20. Choe MM, Sporn PH, Swartz MA. Extracellular matrix remodeling by dynamic strain in a three-dimensional tissue-engineered human airway wall model. Am J Respir Cell Mol Biol. 2006;35:306–13. doi: 10.1165/rcmb.2005-0443OC. [DOI] [PMC free article] [PubMed] [Google Scholar]
  21. Choudhury S, Wilson MR, Goddard ME, O'Dea KP, Takata M. Mechanisms of early pulmonary neutrophil sequestration in ventilator-induced lung injury in mice. Am J Physiol Lung Cell Mol Physiol. 2004;287:L902–10. doi: 10.1152/ajplung.00187.2004. [DOI] [PubMed] [Google Scholar]
  22. Chu EK, Cheng J, Foley JS, Mecham BH, Owen CA, Haley KJ, Mariani TJ, Kohane IS, Tschumperlin DJ, Drazen JM. Induction of the plasminogen activator system by mechanical stimulation of human bronchial epithelial cells. Am J Respir Cell Mol Biol. 2006;35:628–38. doi: 10.1165/rcmb.2006-0040OC. [DOI] [PMC free article] [PubMed] [Google Scholar]
  23. Chu EK, Foley JS, Cheng J, Patel AS, Drazen JM, Tschumperlin DJ. Bronchial epithelial compression regulates epidermal growth factor receptor family ligand expression in an autocrine manner. Am J Respir Cell Mol Biol. 2005;32:373–80. doi: 10.1165/rcmb.2004-0266OC. [DOI] [PubMed] [Google Scholar]
  24. Cloutier M, Maltais F, Piedboeuf B. Increased distension stimulates distal capillary growth as well as expression of specific angiogenesis genes in fetal mouse lungs. Exp Lung Res. 2008;34:101–13. doi: 10.1080/01902140701884331. [DOI] [PubMed] [Google Scholar]
  25. Copland IB, Martinez F, Kavanagh BP, Engelberts D, McKerlie C, Belik J, Post M. High tidal volume ventilation causes different inflammatory responses in newborn versus adult lung. Am J Respir Crit Care Med. 2004;169:739–48. doi: 10.1164/rccm.200310-1417OC. [DOI] [PubMed] [Google Scholar]
  26. Copland IB, Post M. Stretch-activated signaling pathways responsible for early response gene expression in fetal lung epithelial cells. J Cell Physiol. 2007;210:133–43. doi: 10.1002/jcp.20840. [DOI] [PubMed] [Google Scholar]
  27. Copland IB, Reynaud D, Pace-Asciak C, Post M. Mechanotransduction of stretch-induced prostanoid release by fetal lung epithelial cells. Am J Physiol Lung Cell Mol Physiol. 2006;291:L487–95. doi: 10.1152/ajplung.00510.2005. [DOI] [PubMed] [Google Scholar]
  28. Dane DM, Yan X, Tamhane RM, Johnson RL, Jr., Estrera AS, Hogg DC, Hogg RT, Hsia CC. Retinoic acid-induced alveolar cellular growth does not improve function after right pneumonectomy. J Appl Physiol. 2004;96:1090–6. doi: 10.1152/japplphysiol.00900.2002. [DOI] [PubMed] [Google Scholar]
  29. Dhanireddy S, Altemeier WA, Matute-Bello G, O'Mahony DS, Glenny RW, Martin TR, Liles WC. Mechanical ventilation induces inflammation, lung injury, and extra-pulmonary organ dysfunction in experimental pneumonia. Lab Invest. 2006;86:790–9. doi: 10.1038/labinvest.3700440. [DOI] [PubMed] [Google Scholar]
  30. Dolhnikoff M, Mauad T, Ludwig MS. Extracellular matrix and oscillatory mechanics of rat lung parenchyma in bleomycin-induced fibrosis. Am J Respir Crit Care Med. 1999;160:1750–7. doi: 10.1164/ajrccm.160.5.9812040. [DOI] [PubMed] [Google Scholar]
  31. Dolinay T, Kaminski N, Felgendreher M, Kim HP, Reynolds P, Watkins SC, Karp D, Uhlig S, Choi AM. Gene expression profiling of target genes in ventilator-induced lung injury. Physiol Genomics. 2006;26:68–75. doi: 10.1152/physiolgenomics.00110.2005. [DOI] [PubMed] [Google Scholar]
  32. Donaldson SH, Bennett WD, Zeman KL, Knowles MR, Tarran R, Boucher RC. Mucus clearance and lung function in cystic fibrosis with hypertonic saline. N Engl J Med. 2006;354:241–50. doi: 10.1056/NEJMoa043891. [DOI] [PubMed] [Google Scholar]
  33. dos Santos CC, Han B, Andrade CF, Bai X, Uhlig S, Hubmayr R, Tsang M, Lodyga M, Keshavjee S, Slutsky AS, Liu M. DNA microarray analysis of gene expression in alveolar epithelial cells in response to TNFalpha, LPS, and cyclic stretch. Physiol Genomics. 2004;19:331–42. doi: 10.1152/physiolgenomics.00153.2004. [DOI] [PubMed] [Google Scholar]
  34. dos Santos CC, Okutani D, Hu P, Han B, Crimi E, He X, Keshavjee S, Greenwood C, Slutsky AS, Zhang H, Liu M. Differential gene profiling in acute lung injury identifies injury-specific gene expression. Crit Care Med. 2008;36:855–65. doi: 10.1097/CCM.0B013E3181659333. [DOI] [PubMed] [Google Scholar]
  35. Ebihara T, Venkatesan N, Tanaka R, Ludwig MS. Changes in extracellular matrix and tissue viscoelasticity in bleomycin-induced lung fibrosis. Temporal aspects. Am J Respir Crit Care Med. 2000;162:1569–76. doi: 10.1164/ajrccm.162.4.9912011. [DOI] [PubMed] [Google Scholar]
  36. Engler AJ, Griffin MA, Sen S, Bonnemann CG, Sweeney HL, Discher DE. Myotubes differentiate optimally on substrates with tissue-like stiffness: pathological implications for soft or stiff microenvironments. J Cell Biol. 2004;166:877–87. doi: 10.1083/jcb.200405004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  37. Engler AJ, Rehfeldt F, Sen S, Discher DE. Microtissue elasticity: measurements by atomic force microscopy and its influence on cell differentiation. Methods Cell Biol. 2007;83:521–45. doi: 10.1016/S0091-679X(07)83022-6. [DOI] [PubMed] [Google Scholar]
  38. Fehrenbach H, Voswinckel R, Michl V, Mehling T, Fehrenbach A, Seeger W, Nyengaard JR. Neoalveolarisation contributes to compensatory lung growth following pneumonectomy in mice. Eur Respir J. 2008;31:515–22. doi: 10.1183/09031936.00109407. [DOI] [PubMed] [Google Scholar]
  39. Felder E, Siebenbrunner M, Busch T, Fois G, Miklavc P, Walther P, Dietl P. Mechanical strain of alveolar type II cells in culture: changes in the transcellular cytokeratin network and adaptations. Am J Physiol Lung Cell Mol Physiol. 2008;295:L849–57. doi: 10.1152/ajplung.00503.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  40. Fisher JL, Margulies SS. Modeling the effect of stretch and plasma membrane tension on Na+-K+-ATPase activity in alveolar epithelial cells. Am J Physiol Lung Cell Mol Physiol. 2007;292:L40–53. doi: 10.1152/ajplung.00425.2005. [DOI] [PubMed] [Google Scholar]
  41. Fredberg JJ, Kamm RD. Stress transmission in the lung: pathways from organ to molecule. Annu Rev Physiol. 2006;68:507–41. doi: 10.1146/annurev.physiol.68.072304.114110. [DOI] [PubMed] [Google Scholar]
  42. 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:L210–20. doi: 10.1152/ajplung.00332.2003. [DOI] [PubMed] [Google Scholar]
  43. Fust A, LeBellego F, Iozzo RV, Roughley PJ, Ludwig MS. Alterations in lung mechanics in decorin-deficient mice. Am J Physiol Lung Cell Mol Physiol. 2005;288:L159–66. doi: 10.1152/ajplung.00089.2004. [DOI] [PubMed] [Google Scholar]
  44. Gavara N, Roca-Cusachs P, Sunyer R, Farre R, Navajas D. Mapping cell-matrix stresses during stretch reveals inelastic reorganization of the cytoskeleton. Biophys J. 2008;95:464–71. doi: 10.1529/biophysj.107.124180. [DOI] [PMC free article] [PubMed] [Google Scholar]
  45. Gharib SA, Liles WC, Matute-Bello G, Glenny RW, Martin TR, Altemeier WA. Computational identification of key biological modules and transcription factors in acute lung injury. Am J Respir Crit Care Med. 2006;173:653–8. doi: 10.1164/rccm.200509-1473OC. [DOI] [PubMed] [Google Scholar]
  46. Goffin JM, Pittet P, Csucs G, Lussi JW, Meister JJ, Hinz B. Focal adhesion size controls tension-dependent recruitment of alpha-smooth muscle actin to stress fibers. J Cell Biol. 2006;172:259–68. doi: 10.1083/jcb.200506179. [DOI] [PMC free article] [PubMed] [Google Scholar]
  47. Grigoryev DN, Ma SF, Irizarry RA, Ye SQ, Quackenbush J, Garcia JG. Orthologous gene-expression profiling in multi-species models: search for candidate genes. Genome Biol. 2004;5:R34. doi: 10.1186/gb-2004-5-5-r34. [DOI] [PMC free article] [PubMed] [Google Scholar]
  48. Hamanaka K, Jian MY, Weber DS, Alvarez DF, Townsley MI, Al-Mehdi AB, King JA, Liedtke W, Parker JC. TRPV4 initiates the acute calcium-dependent permeability increase during ventilator-induced lung injury in isolated mouse lungs. Am J Physiol Lung Cell Mol Physiol. 2007;293:L923–32. doi: 10.1152/ajplung.00221.2007. [DOI] [PubMed] [Google Scholar]
  49. Hammerschmidt S, Kuhn H, Gessner C, Seyfarth HJ, Wirtz H. Stretch-induced alveolar type II cell apoptosis: role of endogenous bradykinin and PI3K-Akt signaling. Am J Respir Cell Mol Biol. 2007;37:699–705. doi: 10.1165/rcmb.2006-0429OC. [DOI] [PubMed] [Google Scholar]
  50. Hammerschmidt S, Kuhn H, Grasenack T, Gessner C, Wirtz H. Apoptosis and necrosis induced by cyclic mechanical stretching in alveolar type II cells. Am J Respir Cell Mol Biol. 2004;30:396–402. doi: 10.1165/rcmb.2003-0136OC. [DOI] [PubMed] [Google Scholar]
  51. Hammerschmidt S, Kuhn H, Sack U, Schlenska A, Gessner C, Gillissen A, Wirtz H. Mechanical stretch alters alveolar type II cell mediator release toward a proinflammatory pattern. Am J Respir Cell Mol Biol. 2005;33:203–10. doi: 10.1165/rcmb.2005-0067OC. [DOI] [PubMed] [Google Scholar]
  52. Hoganson DM, Pryor HI, 2nd, Vacanti JP. Tissue engineering and organ structure: a vascularized approach to liver and lung. Pediatr Res. 2008;63:520–6. doi: 10.1203/01.pdr.0000305879.38476.0c. [DOI] [PubMed] [Google Scholar]
  53. Hsia CC. Signals and mechanisms of compensatory lung growth. J Appl Physiol. 2004;97:1992–8. doi: 10.1152/japplphysiol.00530.2004. [DOI] [PubMed] [Google Scholar]
  54. Ito S, Bartolak-Suki E, Shipley JM, Parameswaran H, Majumdar A, Suki B. Early emphysema in the tight skin and pallid mice: roles of microfibril-associated glycoproteins, collagen, and mechanical forces. Am J Respir Cell Mol Biol. 2006;34:688–94. doi: 10.1165/rcmb.2006-0002OC. [DOI] [PMC free article] [PubMed] [Google Scholar]
  55. Ito S, Ingenito EP, Arold SP, Parameswaran H, Tgavalekos NT, Lutchen KR, Suki B. Tissue heterogeneity in the mouse lung: effects of elastase treatment. J Appl Physiol. 2004;97:204–12. doi: 10.1152/japplphysiol.01246.2003. [DOI] [PubMed] [Google Scholar]
  56. Ito S, Ingenito EP, Brewer KK, Black LD, Parameswaran H, Lutchen KR, Suki B. Mechanics, nonlinearity, and failure strength of lung tissue in a mouse model of emphysema: possible role of collagen remodeling. J Appl Physiol. 2005;98:503–11. doi: 10.1152/japplphysiol.00590.2004. [DOI] [PubMed] [Google Scholar]
  57. James AL, Wenzel S. Clinical relevance of airway remodelling in airway diseases. Eur Respir J. 2007;30:134–55. doi: 10.1183/09031936.00146905. [DOI] [PubMed] [Google Scholar]
  58. Janes KA, Albeck JG, Gaudet S, Sorger PK, Lauffenburger DA, Yaffe MB. A systems model of signaling identifies a molecular basis set for cytokine-induced apoptosis. Science. 2005;310:1646–53. doi: 10.1126/science.1116598. [DOI] [PubMed] [Google Scholar]
  59. Jesudason R, Black L, Majumdar A, Stone P, Suki B. Differential effects of static and cyclic stretching during elastase digestion on the mechanical properties of extracellular matrices. J Appl Physiol. 2007;103:803–11. doi: 10.1152/japplphysiol.00057.2007. [DOI] [PubMed] [Google Scholar]
  60. Jones JC, Lane K, Hopkinson SB, Lecuona E, Geiger RC, Dean DA, Correa-Meyer E, Gonzales M, Campbell K, Sznajder JI, Budinger S. Laminin-6 assembles into multimolecular fibrillar complexes with perlecan and participates in mechanical-signal transduction via a dystroglycan-dependent, integrin-independent mechanism. J Cell Sci. 2005;118:2557–66. doi: 10.1242/jcs.02395. [DOI] [PMC free article] [PubMed] [Google Scholar]
  61. Kang I, Panneerselvam D, Panoskaltsis VP, Eppell SJ, Marchant RE, Doerschuk CM. Changes in the hyperelastic properties of endothelial cells induced by tumor necrosis factor-alpha. Biophys J. 2008;94:3273–85. doi: 10.1529/biophysj.106.099333. [DOI] [PMC free article] [PubMed] [Google Scholar]
  62. Kay SS, Bilek AM, Dee KC, Gaver DP., 3rd Pressure gradient, not exposure duration, determines the extent of epithelial cell damage in a model of pulmonary airway reopening. J Appl Physiol. 2004;97:269–76. doi: 10.1152/japplphysiol.01288.2003. [DOI] [PubMed] [Google Scholar]
  63. Kim CF. Paving the road for lung stem cell biology: bronchioalveolar stem cells and other putative distal lung stem cells. Am J Physiol Lung Cell Mol Physiol. 2007;293:L1092–8. doi: 10.1152/ajplung.00015.2007. [DOI] [PubMed] [Google Scholar]
  64. Kojic N, Kojic M, Tschumperlin DJ. Computational modeling of extracellular mechanotransduction. Biophys J. 2006;90:4261–70. doi: 10.1529/biophysj.105.078345. [DOI] [PMC free article] [PubMed] [Google Scholar]
  65. Kononov S, Brewer K, Sakai H, Cavalcante FS, Sabayanagam CR, Ingenito EP, Suki B. Roles of mechanical forces and collagen failure in the development of elastase-induced emphysema. Am J Respir Crit Care Med. 2001;164:1920–6. doi: 10.1164/ajrccm.164.10.2101083. [DOI] [PubMed] [Google Scholar]
  66. Kornecki A, Tsuchida S, Ondiveeran HK, Engelberts D, Frndova H, Tanswell AK, Post M, McKerlie C, Belik J, Fox-Robichaud A, Kavanagh BP. Lung development and susceptibility to ventilator-induced lung injury. Am J Respir Crit Care Med. 2005;171:743–52. doi: 10.1164/rccm.200408-1053OC. [DOI] [PubMed] [Google Scholar]
  67. Kuebler WM, Parthasarathi K, Lindert J, Bhattacharya J. Real-time lung microscopy. J Appl Physiol. 2007;102:1255–64. doi: 10.1152/japplphysiol.00786.2006. [DOI] [PubMed] [Google Scholar]
  68. Lee HS, Wang Y, Maciejewski BS, Esho K, Fulton C, Sharma S, Sanchez-Esteban J. Interleukin-10 protects cultured fetal rat type II epithelial cells from injury induced by mechanical stretch. Am J Physiol Lung Cell Mol Physiol. 2008;294:L225–32. doi: 10.1152/ajplung.00370.2007. [DOI] [PubMed] [Google Scholar]
  69. Leung LY, Tian D, Brangwynne CP, Weitz DA, Tschumperlin DJ. A new microrheometric approach reveals individual and cooperative roles for TGF-beta1 and IL-1beta in fibroblast-mediated stiffening of collagen gels. Faseb J. 2007;21:2064–73. doi: 10.1096/fj.06-7510com. [DOI] [PubMed] [Google Scholar]
  70. Levine GK, Deutschman CS, Helfaer MA, Margulies SS. Sepsis-induced lung injury in rats increases alveolar epithelial vulnerability to stretch. Crit Care Med. 2006;34:1746–51. doi: 10.1097/01.CCM.0000218813.77367.E2. [DOI] [PubMed] [Google Scholar]
  71. Li LF, Yu L, Quinn DA. Ventilation-induced neutrophil infiltration depends on c-Jun N-terminal kinase. Am J Respir Crit Care Med. 2004;169:518–24. doi: 10.1164/rccm.200305-660OC. [DOI] [PubMed] [Google Scholar]
  72. Li Z, Dranoff JA, Chan EP, Uemura M, Sevigny J, Wells RG. Transforming growth factor-beta and substrate stiffness regulate portal fibroblast activation in culture. Hepatology. 2007;46:1246–56. doi: 10.1002/hep.21792. [DOI] [PubMed] [Google Scholar]
  73. Ma SF, Grigoryev DN, Taylor AD, Nonas S, Sammani S, Ye SQ, Garcia JG. Bioinformatic identification of novel early stress response genes in rodent models of lung injury. Am J Physiol Lung Cell Mol Physiol. 2005;289:L468–77. doi: 10.1152/ajplung.00109.2005. [DOI] [PubMed] [Google Scholar]
  74. Maina JN, West JB. Thin and strong! The bioengineering dilemma in the structural and functional design of the blood-gas barrier. Physiol Rev. 2005;85:811–44. doi: 10.1152/physrev.00022.2004. [DOI] [PubMed] [Google Scholar]
  75. Mascarenhas MM, Day RM, Ochoa CD, Choi WI, Yu L, Ouyang B, Garg HG, Hales CA, Quinn DA. Low molecular weight hyaluronan from stretched lung enhances interleukin-8 expression. Am J Respir Cell Mol Biol. 2004;30:51–60. doi: 10.1165/rcmb.2002-0167OC. [DOI] [PubMed] [Google Scholar]
  76. McGee KP, Hubmayr RD, Ehman RL. MR elastography of the lung with hyperpolarized 3He. Magn Reson Med. 2008;59:14–8. doi: 10.1002/mrm.21465. [DOI] [PubMed] [Google Scholar]
  77. McParland BE, Macklem PT, Pare PD. Airway wall remodeling: friend or foe? J Appl Physiol. 2003;95:426–34. doi: 10.1152/japplphysiol.00159.2003. [DOI] [PubMed] [Google Scholar]
  78. Meade MO, Cook DJ, Guyatt GH, Slutsky AS, Arabi YM, Cooper DJ, Davies AR, Hand LE, Zhou Q, Thabane L, Austin P, Lapinsky S, Baxter A, Russell J, Skrobik Y, Ronco JJ, Stewart TE. Ventilation strategy using low tidal volumes, recruitment maneuvers, and high positive end-expiratory pressure for acute lung injury and acute respiratory distress syndrome: a randomized controlled trial. Jama. 2008;299:637–45. doi: 10.1001/jama.299.6.637. [DOI] [PubMed] [Google Scholar]
  79. Miller-Jensen K, Janes KA, Brugge JS, Lauffenburger DA. Common effector processing mediates cell-specific responses to stimuli. Nature. 2007;448:604–8. doi: 10.1038/nature06001. [DOI] [PubMed] [Google Scholar]
  80. Miyahara T, Hamanaka K, Weber DS, Drake DA, Anghelescu M, Parker JC. Phosphoinositide 3-kinase, Src, and Akt modulate acute ventilation-induced vascular permeability increases in mouse lungs. Am J Physiol Lung Cell Mol Physiol. 2007;293:L11–21. doi: 10.1152/ajplung.00279.2005. [DOI] [PubMed] [Google Scholar]
  81. Moore KA, Huang S, Kong Y, Sunday ME, Ingber DE. Control of embryonic lung branching morphogenesis by the Rho activator, cytotoxic necrotizing factor 1. J Surg Res. 2002;104:95–100. doi: 10.1006/jsre.2002.6418. [DOI] [PubMed] [Google Scholar]
  82. Moore KA, Polte T, Huang S, Shi B, Alsberg E, Sunday ME, Ingber DE. Control of basement membrane remodeling and epithelial branching morphogenesis in embryonic lung by Rho and cytoskeletal tension. Dev Dyn. 2005;232:268–81. doi: 10.1002/dvdy.20237. [DOI] [PubMed] [Google Scholar]
  83. Nichols JE, Cortiella J. Engineering of a complex organ: progress toward development of a tissue-engineered lung. Proc Am Thorac Soc. 2008;5:723–30. doi: 10.1513/pats.200802-022AW. [DOI] [PubMed] [Google Scholar]
  84. Nolen-Walston RD, Kim CF, Mazan MR, Ingenito EP, Gruntman AM, Tsai L, Boston R, Woolfenden AE, Jacks T, Hoffman AM. Cellular kinetics and modeling of bronchioalveolar stem cell response during lung regeneration. Am J Physiol Lung Cell Mol Physiol. 2008;294:L1158–65. doi: 10.1152/ajplung.00298.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  85. O'Mahony DS, Liles WC, Altemeier WA, Dhanireddy S, Frevert CW, Liggitt D, Martin TR, Matute-Bello G. Mechanical ventilation interacts with endotoxemia to induce extrapulmonary organ dysfunction. Crit Care. 2006;10:R136. doi: 10.1186/cc5050. [DOI] [PMC free article] [PubMed] [Google Scholar]
  86. Obladen M. History of surfactant up to 1980. Biol Neonate. 2005;87:308–16. doi: 10.1159/000084878. [DOI] [PubMed] [Google Scholar]
  87. Oeckler RA, Hubmayr RD. Ventilator-associated lung injury: a search for better therapeutic targets. Eur Respir J. 2007;30:1216–26. doi: 10.1183/09031936.00104907. [DOI] [PubMed] [Google Scholar]
  88. Pan J, Copland I, Post M, Yeger H, Cutz E. Mechanical stretch-induced serotonin release from pulmonary neuroendocrine cells: implications for lung development. Am J Physiol Lung Cell Mol Physiol. 2006;290:L185–93. doi: 10.1152/ajplung.00167.2005. [DOI] [PubMed] [Google Scholar]
  89. Papaiahgari S, Yerrapureddy A, Hassoun PM, Garcia JG, Birukov KG, Reddy SP. EGFR-activated signaling and actin remodeling regulate cyclic stretch-induced NRF2-ARE activation. Am J Respir Cell Mol Biol. 2007;36:304–12. doi: 10.1165/rcmb.2006-0131OC. [DOI] [PMC free article] [PubMed] [Google Scholar]
  90. Park JA, Tschumperlin DJ. Chronic Intermittent Mechanical Stress Increases MUC5AC Protein Expression. Am J Respir Cell Mol Biol. 2009 doi: 10.1165/rcmb.2008-0195OC. [DOI] [PMC free article] [PubMed] [Google Scholar]
  91. Patel AS, Reigada D, Mitchell CH, Bates SR, Margulies SS, Koval M. Paracrine stimulation of surfactant secretion by extracellular ATP in response to mechanical deformation. Am J Physiol Lung Cell Mol Physiol. 2005;289:L489–96. doi: 10.1152/ajplung.00074.2005. [DOI] [PubMed] [Google Scholar]
  92. Perlman CE, Bhattacharya J. Alveolar expansion imaged by optical sectioning microscopy. J Appl Physiol. 2007;103:1037–44. doi: 10.1152/japplphysiol.00160.2007. [DOI] [PubMed] [Google Scholar]
  93. Popp A, Wendel M, Knels L, Koch T, Koch E. Imaging of the three-dimensional alveolar structure and the alveolar mechanics of a ventilated and perfused isolated rabbit lung with Fourier domain optical coherence tomography. J Biomed Opt. 2006;11:014015. doi: 10.1117/1.2162158. [DOI] [PubMed] [Google Scholar]
  94. Pugin J, Dunn-Siegrist I, Dufour J, Tissieres P, Charles PE, Comte R. Cyclic stretch of human lung cells induces an acidification and promotes bacterial growth. Am J Respir Cell Mol Biol. 2008;38:362–70. doi: 10.1165/rcmb.2007-0114OC. [DOI] [PubMed] [Google Scholar]
  95. Raub CB, Suresh V, Krasieva T, Lyubovitsky J, Mih JD, Putnam AJ, Tromberg BJ, George SC. Noninvasive assessment of collagen gel microstructure and mechanics using multiphoton microscopy. Biophys J. 2007;92:2212–22. doi: 10.1529/biophysj.106.097998. [DOI] [PMC free article] [PubMed] [Google Scholar]
  96. Raub CB, Unruh J, Suresh V, Krasieva T, Lindmo T, Gratton E, Tromberg BJ, George SC. Image correlation spectroscopy of multiphoton images correlates with collagen mechanical properties. Biophys J. 2008;94:2361–73. doi: 10.1529/biophysj.107.120006. [DOI] [PMC free article] [PubMed] [Google Scholar]
  97. Ravikumar P, Yilmaz C, Dane DM, Johnson RL, Jr., Estrera AS, Hsia CC. Developmental signals do not further accentuate nonuniform postpneumonectomy compensatory lung growth. J Appl Physiol. 2007;102:1170–7. doi: 10.1152/japplphysiol.00520.2006. [DOI] [PubMed] [Google Scholar]
  98. Rawlins EL, Hogan BL. Epithelial stem cells of the lung: privileged few or opportunities for many? Development. 2006;133:2455–65. doi: 10.1242/dev.02407. [DOI] [PubMed] [Google Scholar]
  99. Ritter MC, Jesudason R, Majumdar A, Stamenovic D, Buczek-Thomas JA, Stone PJ, Nugent MA, Suki B. A zipper network model of the failure mechanics of extracellular matrices. Proc Natl Acad Sci U S A. 2009;106:1081–6. doi: 10.1073/pnas.0808414106. [DOI] [PMC free article] [PubMed] [Google Scholar]
  100. Sabouri-Ghomi M, Wu Y, Hahn K, Danuser G. Visualizing and quantifying adhesive signals. Curr Opin Cell Biol. 2008;20:541–50. doi: 10.1016/j.ceb.2008.05.004. [DOI] [PMC free article] [PubMed] [Google Scholar]
  101. Sanchez-Esteban J, Wang Y, Filardo EJ, Rubin LP, Ingber DE. Integrins beta1, alpha6, and alpha3 contribute to mechanical strain-induced differentiation of fetal lung type II epithelial cells via distinct mechanisms. Am J Physiol Lung Cell Mol Physiol. 2006;290:L343–50. doi: 10.1152/ajplung.00189.2005. [DOI] [PubMed] [Google Scholar]
  102. Sanchez-Esteban J, Wang Y, Gruppuso PA, Rubin LP. Mechanical stretch induces fetal type II cell differentiation via an epidermal growth factor receptor-extracellular-regulated protein kinase signaling pathway. Am J Respir Cell Mol Biol. 2004;30:76–83. doi: 10.1165/rcmb.2003-0121OC. [DOI] [PubMed] [Google Scholar]
  103. Seaborn T, St-Amand J, Cloutier M, Tremblay MG, Maltais F, Dinel S, Moulin V, Khan PA, Piedboeuf B. Identification of cellular processes that are rapidly modulated in response to tracheal occlusion within mice lungs. Pediatr Res. 2008;63:124–30. doi: 10.1203/PDR.0b013e31815eba47. [DOI] [PubMed] [Google Scholar]
  104. Sera T, Fujioka H, Yokota H, Makinouchi A, Himeno R, Schroter RC, Tanishita K. Localized compliance of small airways in excised rat lungs using microfocal X-ray computed tomography. J Appl Physiol. 2004;96:1665–73. doi: 10.1152/japplphysiol.00624.2003. [DOI] [PubMed] [Google Scholar]
  105. Simon BA, Easley RB, Grigoryev DN, Ma SF, Ye SQ, Lavoie T, Tuder RM, Garcia JG. Microarray analysis of regional cellular responses to local mechanical stress in acute lung injury. Am J Physiol Lung Cell Mol Physiol. 2006;291:L851–61. doi: 10.1152/ajplung.00463.2005. [DOI] [PubMed] [Google Scholar]
  106. Simpson KJ, Selfors LM, Bui J, Reynolds A, Leake D, Khvorova A, Brugge JS. Identification of genes that regulate epithelial cell migration using an siRNA screening approach. Nat Cell Biol. 2008;10:1027–38. doi: 10.1038/ncb1762. [DOI] [PubMed] [Google Scholar]
  107. Sinclair SE, Molthen RC, Haworth ST, Dawson CA, Waters CM. Airway strain during mechanical ventilation in an intact animal model. Am J Respir Crit Care Med. 2007;176:786–94. doi: 10.1164/rccm.200701-088OC. [DOI] [PMC free article] [PubMed] [Google Scholar]
  108. Stripp BR, Shapiro SD. Stem cells in lung disease, repair, and the potential for therapeutic interventions: State-of-the-art and future challenges. Am J Respir Cell Mol Biol. 2006;34:517–18. doi: 10.1165/rcmb.F315. [DOI] [PubMed] [Google Scholar]
  109. Suki B, Bates JH. Extracellular matrix mechanics in lung parenchymal diseases. Respir Physiol Neurobiol. 2008;163:33–43. doi: 10.1016/j.resp.2008.03.015. [DOI] [PMC free article] [PubMed] [Google Scholar]
  110. Suki B, Ito S, Stamenovic D, Lutchen KR, Ingenito EP. Biomechanics of the lung parenchyma: critical roles of collagen and mechanical forces. J Appl Physiol. 2005;98:1892–9. doi: 10.1152/japplphysiol.01087.2004. [DOI] [PubMed] [Google Scholar]
  111. Swartz MA, Tschumperlin DJ, Kamm RD, Drazen JM. Mechanical stress is communicated between different cell types to elicit matrix remodeling. Proc Natl Acad Sci U S A. 2001;98:6180–5. doi: 10.1073/pnas.111133298. [DOI] [PMC free article] [PubMed] [Google Scholar]
  112. Tarran R, Button B, Boucher RC. Regulation of normal and cystic fibrosis airway surface liquid volume by phasic shear stress. Annu Rev Physiol. 2006;68:543–61. doi: 10.1146/annurev.physiol.68.072304.112754. [DOI] [PubMed] [Google Scholar]
  113. Tarran R, Button B, Picher M, Paradiso AM, Ribeiro CM, Lazarowski ER, Zhang L, Collins PL, Pickles RJ, Fredberg JJ, Boucher RC. Normal and cystic fibrosis airway surface liquid homeostasis. The effects of phasic shear stress and viral infections. J Biol Chem. 2005;280:35751–9. doi: 10.1074/jbc.M505832200. [DOI] [PMC free article] [PubMed] [Google Scholar]
  114. Tarran R, Donaldson S, Boucher RC. Rationale for hypertonic saline therapy for cystic fibrosis lung disease. Semin Respir Crit Care Med. 2007;28:295–302. doi: 10.1055/s-2007-981650. [DOI] [PubMed] [Google Scholar]
  115. Tawhai MH, Burrowes KS. Multi-scale models of the lung airways and vascular system. Adv Exp Med Biol. 2008;605:190–4. doi: 10.1007/978-0-387-73693-8_33. [DOI] [PubMed] [Google Scholar]
  116. Trepat X, Deng L, An SS, Navajas D, Tschumperlin DJ, Gerthoffer WT, Butler JP, Fredberg JJ. Universal physical responses to stretch in the living cell. Nature. 2007;447:592–5. doi: 10.1038/nature05824. [DOI] [PMC free article] [PubMed] [Google Scholar]
  117. Tschumperlin DJ, Dai G, Maly IV, Kikuchi T, Laiho LH, McVittie AK, Haley KJ, Lilly CM, So PT, Lauffenburger DA, Kamm RD, Drazen JM. Mechanotransduction through growth-factor shedding into the extracellular space. Nature. 2004;429:83–6. doi: 10.1038/nature02543. [DOI] [PMC free article] [PubMed] [Google Scholar]
  118. Tschumperlin DJ, Drazen JM. Chronic effects of mechanical force on airways. Annu Rev Physiol. 2006;68:563–83. doi: 10.1146/annurev.physiol.68.072304.113102. [DOI] [PubMed] [Google Scholar]
  119. Tschumperlin DJ, Margulies SS. Equibiaxial deformation-induced injury of alveolar epithelial cells in vitro. Am J Physiol. 1998;275:L1173–83. doi: 10.1152/ajplung.1998.275.6.L1173. [DOI] [PubMed] [Google Scholar]
  120. Tsuchida S, Engelberts D, Peltekova V, Hopkins N, Frndova H, Babyn P, McKerlie C, Post M, McLoughlin P, Kavanagh BP. Atelectasis causes alveolar injury in nonatelectatic lung regions. Am J Respir Crit Care Med. 2006;174:279–89. doi: 10.1164/rccm.200506-1006OC. [DOI] [PubMed] [Google Scholar]
  121. Tsuchida S, Engelberts D, Roth M, McKerlie C, Post M, Kavanagh BP. Continuous positive airway pressure causes lung injury in a model of sepsis. Am J Physiol Lung Cell Mol Physiol. 2005;289:L554–64. doi: 10.1152/ajplung.00143.2005. [DOI] [PubMed] [Google Scholar]
  122. Unbekandt M, del Moral PM, Sala FG, Bellusci S, Warburton D, Fleury V. Tracheal occlusion increases the rate of epithelial branching of embryonic mouse lung via the FGF10-FGFR2b-Sprouty2 pathway. Mech Dev. 2008;125:314–24. doi: 10.1016/j.mod.2007.10.013. [DOI] [PMC free article] [PubMed] [Google Scholar]
  123. Vlahakis NE, Hubmayr RD. Cellular stress failure in ventilator-injured lungs. Am J Respir Crit Care Med. 2005;171:1328–42. doi: 10.1164/rccm.200408-1036SO. [DOI] [PMC free article] [PubMed] [Google Scholar]
  124. Wagers S, Lundblad LK, Ekman M, Irvin CG, Bates JH. The allergic mouse model of asthma: normal smooth muscle in an abnormal lung? J Appl Physiol. 2004;96:2019–27. doi: 10.1152/japplphysiol.00924.2003. [DOI] [PubMed] [Google Scholar]
  125. Wagers SS, Haverkamp HC, Bates JH, Norton RJ, Thompson-Figueroa JA, Sullivan MJ, Irvin CG. Intrinsic and antigen-induced airway hyperresponsiveness are the result of diverse physiological mechanisms. J Appl Physiol. 2007;102:221–30. doi: 10.1152/japplphysiol.01385.2005. [DOI] [PubMed] [Google Scholar]
  126. Wagh AA, Roan E, Chapman KE, Desai LP, Rendon DA, Eckstein EC, Waters CM. Localized elasticity measured in epithelial cells migrating at a wound edge using atomic force microscopy. Am J Physiol Lung Cell Mol Physiol. 2008;295:L54–60. doi: 10.1152/ajplung.00475.2007. [DOI] [PMC free article] [PubMed] [Google Scholar]
  127. Wang HB, Dembo M, Wang YL. Substrate flexibility regulates growth and apoptosis of normal but not transformed cells. Am J Physiol Cell Physiol. 2000;279:C1345–50. doi: 10.1152/ajpcell.2000.279.5.C1345. [DOI] [PubMed] [Google Scholar]
  128. Wang Y, Maciejewski BS, Lee N, Silbert O, McKnight NL, Frangos JA, Sanchez-Esteban J. Strain-induced fetal type II epithelial cell differentiation is mediated via cAMP-PKA-dependent signaling pathway. Am J Physiol Lung Cell Mol Physiol. 2006;291:L820–7. doi: 10.1152/ajplung.00068.2006. [DOI] [PubMed] [Google Scholar]
  129. Wang Y, Maciejewski BS, Weissmann G, Silbert O, Han H, Sanchez-Esteban J. DNA microarray reveals novel genes induced by mechanical forces in fetal lung type II epithelial cells. Pediatr Res. 2006;60:118–24. doi: 10.1203/01.pdr.0000227479.73003.b5. [DOI] [PubMed] [Google Scholar]
  130. Wipff PJ, Rifkin DB, Meister JJ, Hinz B. Myofibroblast contraction activates latent TGF-beta1 from the extracellular matrix. J Cell Biol. 2007;179:1311–23. doi: 10.1083/jcb.200704042. [DOI] [PMC free article] [PubMed] [Google Scholar]
  131. Wurfel MM. Microarray-based analysis of ventilator-induced lung injury. Proc Am Thorac Soc. 2007;4:77–84. doi: 10.1513/pats.200608-149JG. [DOI] [PMC free article] [PubMed] [Google Scholar]
  132. Yalcin HC, Perry SF, Ghadiali SN. Influence of airway diameter and cell confluence on epithelial cell injury in an in vitro model of airway reopening. J Appl Physiol. 2007;103:1796–807. doi: 10.1152/japplphysiol.00164.2007. [DOI] [PubMed] [Google Scholar]
  133. Zani BG, Kojima K, Vacanti CA, Edelman ER. Tissue-engineered endothelial and epithelial implants differentially and synergistically regulate airway repair. Proc Natl Acad Sci U S A. 2008;105:7046–51. doi: 10.1073/pnas.0802463105. [DOI] [PMC free article] [PubMed] [Google Scholar]

RESOURCES