Evaluation of lung injury in rats and mice
James C. Parker and
Mary I. Townsley
Department of Physiology, University of South Alabama, Mobile, Alabama 36688
 |
ABSTRACT
|
---|
Lung injury is a broad descriptor that can be applied to conditions ranging from mild interstitial edema without cellular injury to massive and fatal destruction of the lung. This review addresses those methods that can be readily applied to rats and mice whose small size limits the techniques that can be practically used to assess injury. The methodologies employed range from nonspecific measurement of edema formation to techniques for calculating values of specific permeability coefficient for the microvascular membrane in lung. Accumulation of pulmonary edema can be easily and quantitatively measured using gravimetric methods and indicates an imbalance in filtration forces or restrictive properties of the microvascular barrier. Lung compliance can be continuously measured, and light and electron microscopy can be used regardless of lung size to detect edema and structural damage. Increases in fluid and/or protein flux due to increased permeability must also be separated from those due to increased filtration pressure for mechanistic interpretation. Although an increase in the initial lung albumin clearance compared with controls matched for size and filtration pressure is a reliable indicator of endothelial dysfunction, calculated alterations in capillary filtration coefficient Kf,c, reflection coefficient
, and permeability-surface area product PS are the most accurate indicators of increased permeability. Generally, PS and Kf,c will increase and
will decrease with vascular injury, but derecruitment of microvascular surface area may attenuate the affect on PS and Kf,c without altering measurements of
.
edema; acute respiratory distress syndrome
IN THE LUNG, FLUID AND SOLUTE FLUX into the adjacent interstitial space are regulated by net transvascular gradients for hydrostatic forces and protein concentrations as well as by permeability of the pulmonary endothelial barrier to water and solutes. Fluid filtered from the vascular space into the interstitium then percolates through the interstitium to enter the initial lymphatics. Interstitial edema forms when an imbalance exists between the rate of fluid filtration (Jv) into the pulmonary interstitium and the rate of fluid exit via the lymphatic system (7, 118). Accumulation of pulmonary edema can result when either the net transcapillary filtration pressure or the permeability of the pulmonary microvascular endothelial barrier increases, such as in acute or chronic pulmonary hypertension or in acute respiratory distress syndrome (ARDS). However, these two mechanisms are often difficult to separate, particularly in vivo in small animals such as mice, where pulmonary hemodynamics are difficult to measure. Although the basic theory underlying transvascular fluid and protein movement is certainly applicable regardless of lung size, resolution of the mechanisms causing edema formation and lung injury in rats, and especially in mice, poses a considerable technical challenge for the investigator. This review presents methods that can be applied in both rats and mice, as well as larger species, to evaluate mechanisms of lung edema secondary to pulmonary vascular hypertension and endothelial dysfunction.
 |
THEORY
|
---|
It is important to understand the interrelationship between transvascular fluid and protein flux (Js) because this poses the underlying difficulty in differentiating between hydrostatic and permeability edema in intact lungs. The rate of transvascular Jv is a function of the net balance of hydrostatic and colloid osmotic forces across the capillary barrier and capillary endothelial integrity (i.e., microvascular permeability)
 | (1) |
Jv has traditionally been defined as a process occurring in the capillary network in the lung, where Jv is determined by the "Starling forces" acting across the endothelial barrier. These forces include the pulmonary capillary pressure (Pc), interstitial fluid pressure (Pt), and the plasma and interstitial colloid osmotic pressures (
p and
t, respectively), where the interstitial forces are those existing in the interstitium of the alveolar septal wall and perivascular spaces around small extra-alveolar vessels. However, recent theoretical and experimental studies suggest that the effective colloid osmotic pressure of tissue proteins is somewhat less than that predicted by the average tissue protein concentration due to relatively high local flow velocities at the sieving pores, even when net filtration is low (55). During alveolar flooding, the surface tension generated at the fluid interface of partially filled alveoli can create an additional significant driving force for further alveolar filling (167). The capillary filtration coefficient (Kf,c) is the product of the endothelial hydraulic conductance (Lp) and endothelial surface area (S) available for filtration. Kf,c and the osmotic reflection coefficient (
), which describes the endothelial selectivity for total proteins, dictate the effectiveness of transcapillary forces in producing fluid and protein movement across the capillaries (7, 118). Because Kf,c is proportional to the amount of perfused exchange area in the lung (177), transcapillary filtration rate can vary significantly as vascular area is recruited or derecruited, even when the net Starling force balance remains constant.
Although it may not be immediately obvious, transvascular fluid and Js are interrelated because Js is determined both by Jv, which carries proteins convectively across the exchange barrier, and by dissipative processes, such as diffusion or transcytosis of protein
 | (2) |
This relationship was first derived by Patlak et al. (125), later modified by Granger and Taylor (43), and discussed by Rippe and Haraldsson (140). The undefined variables in Eq. 2 are the plasma (Cp) and interstitial fluid (Ct) protein concentrations and the Peclet number, Pe. Pe is a dimensionless number describing the contribution of convective protein flux relative to that occurring via diffusion, i.e., Pe = Jv (1 -
)/PS. The term PS contained within Pe is the permeability-surface area product (PS), where P represents the diffusive permeability for a macromolecule (the third permeability coefficient) and S is the surface area for solute exchange. Both the convective and diffusive components of protein flux likely contribute significantly at normal, resting pulmonary capillary pressures. However, with acute pulmonary venous hypertension and the attendant increase in transvascular fluid flux (i.e., when Jv and Pe increase by >3-fold from baseline), the contribution of diffusive exchange decreases and convective protein flux dominates. Reed et al. (137) have presented a thorough review of this issue. In addition to these passive exchange processes for proteins, some investigators argue that active vesicular transport contributes to movement of proteins across the endothelial barrier (101, 138, 149). Whether vesicular transport contributes significantly to Js at very low filtration rates is still debated, but the preponderance of evidence supports the notion that convective transport of protein is the major mechanism of transport during measurable increases in Jv. Furthermore, Rippe and Taylor (143) recently demonstrated that transcapillary albumin clearances in rat lung were not inhibited by either cooling to prevent active transport or pharmacological blockade of transcytosis, observations that suggest that transcytosis only has a minor role in net transcapillary efflux of albumin.
Vascular permeability to proteins in lungs of small animals is best studied as the unidirectional protein clearance (Clr) because interstitial and lymph protein concentrations cannot be easily sampled. Considering that transcapillary Js can be expressed as Clr where Clr = Js/Cp and that Ct = 0 when a tracer protein is first introduced into the circulation, Eq. 2 for a tracer protein reduces to
 | (3) |
At high filtration rates when Pe exceeds 3, Eq. 3 further reduces to Clr = Jv (1 -
), reflecting the fact that convective transport of protein predominates. When there is no net filtration (Jv = 0), dissipative processes such as diffusion, volume circulation, and/or transcytosis will predominate, and then Clr approximates PS.
The functional implications of interrelationships between the Starling forces, transvascular fluid and Js, and endothelial permeability can be summarized as follows. Increased Pc will not only increase Jv (Eq. 1) and the likelihood of edema formation but will also increase Js (Eq. 2) into the pulmonary interstitium. Similarly, if endothelial permeability increases, Kf,c rises and
falls. A decrease in
contributes to an increase in the effective net filtration pressure. Thus when endothelial permeability increases, the impact on these parameters collectively leads to an increase in Jv (Eq. 1). The injury-induced increase in both Jv and PS, along with the decrease in
, will result in increased Js into the lung interstitium (Eq. 2). Thus the mechanisms responsible for pulmonary edema cannot be fully understood unless specific permeability coefficients are measured.
 |
TECHNIQUES
|
---|
In general, we have grouped the techniques used to evaluate pulmonary edema and endothelial permeability according to their level of difficulty and specificity. In addition to the technological requirements and limitations of each technique, we also discuss particular issues relevant to their use in rats and mice. All of the techniques discussed are applicable to use in larger mammals.
Measurement of Pulmonary Edema
Accumulation of edema fluid first begins in loose perivascular and/or peribronchiolar cuffs, followed by accumulation in the septal interstitium. Alveolar flooding occurs only after extravascular lung water increases by more than
35% (167) and does not occur in a homogenous fashion throughout the lung (195). The pleural space, which acts as a fluid overflow compartment (193), appears to be recruited only when pulmonary lymphatic capacity is overwhelmed and alveolar flooding has already occurred. Hydrostatic edema is specifically the result of a net imbalance in the Starling forces operating across the pulmonary capillary endothelial barrier, generally due to a change in Pc. However, as previously noted, acute lung injury may alter Jv and promote edema formation in the absence of any change in pulmonary vascular hydrostatic pressure. Furthermore, many forms of lung injury are complicated by concomitant pulmonary hypertension, leading to an increase in the net hydrostatic filtration force across the pulmonary exchange barrier and an amplification of edema formation.
Gravimetric assessment of pulmonary edema. The simplest way to evaluate edema formation in lung is to use a gravimetric approach. There are four measures commonly applied: the lung wet weight (LW), the lung wet weight/body weight ratio (LW/BW), the lung wet/dry weight ratio (W/D), and the extravascular lung water (EVLW). Any of these methods is easily applicable to mouse lung, as outlined in Table 1. In general, these techniques require only a sensitive balance and a drying oven. The investigator needs no particular surgical skill. LW and LW/BW ratio have been used in mouse lung to evaluate acute lung injury and the response to dietary supplementation with butylated hydroxyanisole to augment lung antioxidant defense (9, 156). However, comparison of these data with the effect of pulmonary hypertension in chronic congestive heart failure, where LW/BW increased 2.5-fold (90), highlights the difficulty in interpreting these and other gravimetric measures of pulmonary edema. Edema can arise from hydrostatic (i.e., filtration-induced) or injury (increased permeability) mechanisms. Nonetheless, these measures are useful. Because inbred strains have relatively uniform LW/BW ratios between animals at any given age, one can potentially compare LW/BW in treatment groups to that predicted based on body weight. However, the measure of W/D is a more useful tool since it accounts for changes in lung dry mass as well. The W/D ratio had an excellent correlation with bronchoalveolar lavage fluid (BALF) albumin and total protein during graded injury, high-airway pressure lung injury in mice (199). Depending on the experimental paradigm, this may be an important consideration. There is significant variation in the fibrotic potential between some mouse strains (78), which could bias the LW/BW ratio and limit comparisons between strains. To determine W/D, the whole lung, lobes, or segments of peripheral lung are weighed on excision from the animal and then dried in an oven (60-65°C) to constant weight (2-4 days, depending on the initial mass of lung and the lung water content). Interpretation of LW, LW/BW, and W/D can be complicated by the inclusion of blood in the wet lung weight, both from residual intravascular blood and that introduced into the lung interstitium via bleeding or injury. In addition, with any increase in permeability, extravasated protein can contribute to total LW. In either case, these gravimetric measures may reflect significant error. Thus the gravimetric measurement of EVLW, first described by Pearce et al. (128) and modified for use in mouse lung (37), is an improvement due to correction of the lung water content for blood water. To evaluate EVLW, the lung tissue is homogenized with a known aliquot of water, and the aliquots of homogenate, blood, and homogenate supernatant are weighed and then dried to constant weight. LW is corrected for blood weight by comparison with hemoglobin concentration or a radiolabeled tracer in blood and homogenate. Kobayashi et al. (77) used the measure of EVLW to show the impact of inducible nitric oxide synthase (iNOS) in modulating the lung injury response to hyperoxia in mice. They found that EVLW increased from 5.1 to 6.4 ml/g with hyperoxia (see Table 1), but that in mice deficient for iNOS, EVLW increased to 7.9 ml/g after hyperoxic exposure. The additional technical requirements for EVLW include a homogenizer and either a spectrophotometer or a gamma counter. One can also obtain an estimate of residual blood per gram of blood-free dry mass from the EVLW measurement. Parker and Ivey (114) used the measure of residual blood mass in rat lung as one index of extravasation during barotrauma. Although it is more time consuming, EVLW does offer advantages that justify the additional effort.
Other general considerations in using gravimetric approaches for measurement of lung edema in mice are as follows. First, as the mass of the lung tissue to be measured decreases, one must be increasingly careful to weigh the lung as quickly as possible to minimize evaporative losses. Second, investigators should be aware of the impact of ventilation on alveolar fluid clearance in mouse lung. Fukada et al. (37) found that interstitial fluid volume rose and alveolar fluid clearance slowed with time in statically inflated in situ mouse lung compared with that in intact mice when the lung was ventilated. Their observation highlights the importance of consistent treatment of the lungs when such gravimetric measures of edema are utilized. Third, in these small lungs, gravimetric measures yield an assessment of average or "lumped" lung water content and will not reveal any regional heterogeneity. The final and most critical consideration as noted previously is that one cannot differentiate between hydrostatic edema and that due to increased lung endothelial permeability based on these gravimetric measures of edema alone. Concomitant measurement of pulmonary vascular pressure can be useful to document pulmonary hemodynamic status, although this is technically challenging in mice. Champion et al. (18) have developed a fluoroscopic catheterization technique for mice that allows measurement of pulmonary arterial pressure. In addition, the method recently described for measurement of pulmonary arterial pressure in rats via transthoracic echocardiography (61) may also be applicable in mice since echocardiography is utilized in this species (133). The bottom line is that in the absence of specific measures of pulmonary hemodynamics or lung endothelial permeability, the contributions of hydrostatic pressure and vascular permeability to edema formation cannot be reliably separated.
Dynamic and static lung compliance. Respiratory system compliance (Crs) has been used as a nonspecific indicator of acute lung injury, since accumulation of interstitial and alveolar edema and exudate will decrease lung compliance as gas in alveoli and small airways becomes displaced with fluid. Measurement of Crs is useful because it is relatively noninvasive, and it can be used to continuously monitor the progress of lung injury and edema. Dynamic lung compliance can also be measured noninvasively in awake animals using whole body plethysmography (42, 46, 92, 104), although the dynamic measurement of Crs may not be as closely correlated with lung edema as static Crs (2). Sibilla et al. (159) ventilated rats with a wide range of tidal volumes and FIO2 selected to produce ventilator-induced lung injury within 20 min to 5 h. They reported an excellent correlation of Crs with lung W/D and histological scores for lung injury and edema. Crs decreased in proportion to the increase in W/D regardless of the rate of edema formation. Ewart et al. (32) described serial measurements of Crs in mice using a stop-flow method that partitions airway pressure drops due to Crs and airway resistance. This measure has proved to be particularly useful in mice since Crs essentially equals the lung compliance (3.6 ml·cmH2O-1·kg body wt-1) due to the very high chest wall compliance in mice (83, 163). Comparisons with strain-matched control animals are particularly important. Tankersley et al. (163) found that baseline Crs differed significantly among strains of mice, i.e., Crs in C3H/HeJ mice was
2-fold greater than that in C57BL/6J and B6C3F1/J mice.
Noninvasive imaging techniques. A number of noninvasive imaging techniques are available, including X-ray, high-resolution computed tomography (CT), magnetic resonance imaging (MRI), and positron emission tomography (PET). Flat chest X-ray films in larger animals and humans allow assessment of the severity and distribution of edema, based on defined scoring systems that take into account the distribution of edema, pleural effusion, interstitial changes such as septal lines and peribronchial cuffing, and the distribution of pulmonary blood flow and volume (1, 100). The resolution offered by this approach suggests little utility in rats and mice. However, other noninvasive imaging techniques offer potential advantages with respect to the ability to construct three-dimensional images with improved resolution down to 50 µm in some systems. Furthermore, the development of micro-CT and micro-PET bring the scale of these techniques into the range for use in mice. MRI has been extensively used for soft tissue imaging and evaluation of anatomic structure. With the addition of contrast agents, CT also can be applied to soft tissue such as lung. PET is a nuclear imaging technique that can be focused at a particular molecular target and requires use of a positron-emitting radionuclide. These techniques, their limitations, and their applicability to the experimental setting have recently been discussed in detail (11, 19, 86, 127, 130). All three of these techniques have been used in large animals (e.g., dogs) to document regional heterogeneity in the distribution of lung water after lung injury (17, 154, 155, 180). With the advent of small scale instrumentation suitable for rats and mice, high-resolution noninvasive imaging is now used extensively in these animals. The advantage of noninvasive imaging is that one can make sequential measurements over time and that the animal is available for other measurements after the imaging is completed. Neither of these advantages are afforded by gravimetric measures. Imaging does require that the animal be restrained during the procedure, commonly via anesthesia. Furthermore, in rodents, motion artifacts associated with the high cardiac and respiratory rates of the animal (127) limit the resolution of early pulmonary edema. Beckmann et al. (11) used a short echo time and gradient echo images in MRI to minimize these artifacts in rat and mouse lung. With this approach, they found a significant correlation between the signal decay time (T2*) and lung edema assessed by image analysis in spontaneously breathing rats challenged with oleic acid or lipopolysaccharide. These imaging techniques have been used to evaluate lung tumor size and density (65, 127) and lung volume (103) in mice and lung vascular morphometry in rats (63). In our view, the use of imaging techniques to evaluate lung water in mice is probably not warranted due to limitations in resolution. More informative data could be generated by applying these techniques to the three-dimensional evaluation of receptor-ligand interactions or gene expression in lung leading to or resulting from lung injury (15, 60, 130).
Microscopy and Histological Techniques
It should be clear that quantitation of edema fluid alone does not allow discrimination between hydrostatic and permeability edema. Microscopy to evaluate lung structure can be enlightening and can be applied regardless of lung size.
Light and transmission electron microscopy. Light microscopy of lung can provide qualitative evidence supportive of lung injury and/or hydrostatic edema such as alveolar flooding, thickening of alveolar septa, and perivascular or airway cuffing. If the lung is fixed by either vascular perfusion or immersion, accumulation of red blood cells, inflammatory cells, and proteinaceous fluid in the alveolar spaces can be easily detected. Seibert et al. (157) used such an approach with light microscopy of hematoxylin and eosin-stained lung sections to show the qualitative, but clear-cut, beneficial effect of isoproterenol or forskolin in limiting the morphological changes associated with lung ischemic-reperfusion injury. More detailed evaluation of structural changes in lung with injury requires electron microscopy (EM). Transmission EM can be used to elucidate rupture or blebbing of the endothelial barrier and detachment of epithelium and/or endothelium to the basement membrane, as shown in Fig. 1 (8, 30, 189), as well as changes in the character of the endothelial tight junctions. An additional advantage of transmission EM is that one can obtain useful information regarding the heterogeneity of lung injury. Differential sensitivity of the alveolar epithelial and endothelial barriers and the tendency for some types of injury to target the macrovascular vs. microvascular segments of the vasculature have been documented (8, 20, 30). Bachofen et al. (8) found that moderate hydrostatic edema resulted in blebbing and rupture of the epithelial layer in the alveolo-capillary barrier, whereas lesions to the endothelium were rare. Furthermore, they found that barrier lesions were only seen in areas of the lung with alveolar edema, despite the exposure of the entire vasculature to the hydrostatic stress. Chetham et al. (20) found that calcium store depletion-induced injury in rat lung resulted in endothelial gap formation only in vessels
100 µm in diameter. The use of a "lung injury score" based on predetermined criteria and applied to images obtained either by light microscopy or transmission EM has proved useful as a semiquantitative approach to evaluating lung injury (9, 22, 57, 157). Some investigators have taken microscopic evaluation of lung injury one step further and have quantitated structural changes by determining the number of breaks in the endothelium and epithelium per unit barrier length (30) or the number of detachments per capillary (153). For example, Schubert et al. (153) found that in caveolin-1 (-/-) knockout mice, the number of endothelial detachments in lung capillaries increased dramatically from 0.12 per capillary in wild-type mice to >2 per capillary in caveolin-deficient mice. With the use of stereological methods to evaluate images obtained via transmission EM, one can make further quantitative assessment of lung structure relevant to lung injury and edema, including interstitial thickness and the volume fractions of lung occupied by interstitial cells and noncellular interstitium (9, 41, 98, 147, 174, 191).

View larger version (145K):
[in this window]
[in a new window]
|
Fig. 1. Transmission electron micrograph showing a small (30-µm-diameter) vessel in mouse lung. The lung was ventilated for 1 h at 45 cmH2O peak inflation pressure. Note the separation of endothelium and epithelium from the basement membrane (BM) and the accumulation of proteinaceous fluid in the alveolar space. PMN, polymorphonuclear leukocyte.
|
|
Scanning microscopy and corrosion casting. Scanning EM of lung tissue has been used to show rupture of the alveolar epithelium with acute pulmonary hypertension (23). Scanning EM coupled with vascular corrosion casting offers another approach to evaluation of lung structure in normal and injured lung. Either commercially available prepolymerized methyl methacrylate (Mercox) or methyl methacrylate prepolymerized by the investigator using ultraviolet light was applied (38, 152). The lung vasculature is flushed clear of blood, and then the casting material (prepolymerized methyl methacrylate mixed with initiators to facilitate full polymerization) is introduced by pump perfusion. When the cast is fully polymerized, the lung is immersed in soapy water at 55°C overnight to promote tissue maceration, and then corrosion is completed by alternate immersion in 5 M HCl and 5 M NaOH. For rats and mice, corrosion occurs within a week. The cast is washed extensively in distilled water, dried, and fractured. Cast fragments are coated with gold palladium to promote contrast and then visualized using scanning EM. The use of a relatively low viscosity polymer allows the investigator to track or "capture" leak sites induced by lung injury. Parker and colleagues (111a) used low viscosity methyl methacrylate to identify leak sites induced by high-pressure ventilation in mouse lung, as shown in Fig. 2. Townsley et al. (172, 173) have used this approach to show that different injury paradigms have unique targets and patterns of leak within the pulmonary circulation: high vascular pressure resulted in large, but infrequent, leak sites within the alveolar septum, whereas oleic acid injury resulted in numerous, small leaks across the alveolar septal surface.

View larger version (88K):
[in this window]
[in a new window]
|
Fig. 2. Scanning electron micrograph of a vascular corrosion cast of mouse lungs in unventilated mice (A) and mice ventilated for 30 min at a peak inflation pressure of 45 cmH2O (B) (89). Note protrusions of casting material through endothelial leak sites in the injured lung. Leak of casting material was not seen in the cast of unventilated lung.
|
|
Permeability Measurement in Isolated Lungs
Isolated perfused lung preparations have been a standard tool for assessment of lung vascular injury for more than four decades (167). By monitoring vascular and airway pressures as well as lung weight, small changes in vascular resistance and transvascular filtration can be detected and edema formation due to increased vascular filtration pressure can be reliably separated from that due to increased vascular permeability (28). Separation of these effects is critical for determining the mechanisms of edema formation. Isolated, perfused lung preparations can also be used to evaluate inflammatory cell effects and lung-specific production of proinflammatory cytokines (105, 129, 186, 197). The basic equipment required includes a physiological recorder, pressure and weight transducers, a ventilator, a perfusion pump, an arterial bubble trap, and a water bath for temperature control. In small animals, the heart and lungs are usually excised en bloc, and the pulmonary artery and left atrium are cannulated through the right and left ventricles, respectively. The lungs are ventilated with a 5% CO2 gas mixture at the appropriate tidal volume and rate for each species (2 ml at 35 min-1 for rats and 0.2 ml at 120 min-1 for mice). Lungs are perfused with blood or physiological salt solutions containing a colloid such as 4-5% BSA. Addition of autologous or heterologous serum is beneficial, as some serum factors are important for maintaining normal endothelial permeability (48). Lung weight is monitored using a strain gauge force transducer; sensitivity to changes in lung weight can be increased by using a cantilevered beam balance (112).
Filtration coefficients and hydraulic conductivity. The gravimetric Kf,c is a measure of fluid conductance, and as described previously, is the product of Lp and S. Because S is not easily measured in intact lungs, Kf,c is referenced to initial LW or dry lung weight (67, 167). Initial lung weight can be predicted from body weight in inbred species with some degree of accuracy due to uniform body phenotypes (112, 115). When Kf,c is known, Lp can be calculated using estimates of micro-vascular S (167). Gravimetric Kf,c measurements can be repeated several times over 1-6 h, allowing paired statistical comparisons of baseline and experimental states (see Table 2), and have been used extensively to detect injury in isolated rat lungs (69, 105, 109, 115, 116, 178, 197). Despite more challenging surgical techniques, Kf,c has recently been used to evaluate vascular permeability in isolated mouse lungs in the presence of mechanical injury (112), adhesion molecule inhibition with sepsis (40), vascular endothelial-cadherin disassembly by low calcium (39), thrombin receptor knockout mice (184), and transient receptor potential channel knockout mice (171).
Gravimetric Kf,c is measured by comparing the rate of lung weight gain (
W/
t) during the baseline period with that after increasing the venous pressure (Ppv), as shown in Fig. 3. The increase in capillary filtration pressure (
Ppc) is determined by the change in the double occlusion pressure measured before and after the increase in Ppv (175). The increment in Ppc should exceed 5 cmH2O (e.g., 7-10 cmH2O) to assure a filtration rate that can be easily measured (122). Gravimetric Kf,c is calculated by equating filtration rate with the
W/
t measured after the Ppv increase (Eq. 4). The
W/
t can be estimated by logarithmic extrapolation to time 0 (28), curve fitting (67), or simply measuring the weight gain slope (tangent) at some time up to 20 min after the vascular pressure increase (112, 115) where
 | (4) |
If an isogravimetric state is not obtained at baseline, then the baseline
W/
t must be subtracted from the
W/
t induced by increasing the Ppv. Because
W/
t decreases with time after elevating vascular pressure, the time at which the
W/
t is measured is critical, as shown by the change in slopes of the tangents in Fig. 3. The inset to Fig. 3 shows how logarithmic extrapolation of different portions of the lung weight gain curve can lead to different zero intercept values, depending on which portion of the curve is selected. Regression line 1 was fit to log
W/
t between 3 and 10 min, whereas regression line 2 was fit to 18-20 min and indicates that estimates of Kf,c will decrease as later time periods are used for logarithmic curve fitting. Kf,c calculated using the slope at 20 min, i.e., (
W/
t)t=20, has proved more sensitive for detecting increases in transvascular filtration in rodent lung than the slope at earlier times because of a relatively high vascular compliance and a prolonged stress relaxation of pressurized pulmonary vessels (112, 119). Simultaneous comparisons of vascular filtration after an increase in vascular pressure using gravimetric and colorimetric methods based on red blood cell or albumin concentration showed that vascular stress relaxation persisted for
23 min (47, 50, 119). As shown in Table 2, use of the (
W/
t)t=20 slope method results in baseline Kf,c values
40% of those calculated using the time 0 logarithmic extrapolation of the
W/
t data obtained at 3-10 min (3, 28, 36). Kf,c calculated from the logarithmic extrapolation of the slope to time 0, i.e., (
W/
t)t=0, will approach that determined from the (
W/
t)t=20 slope as later time points in the weight gain curve are included in the analysis. This is evident from Fig. 3, inset.
Evaluation of segmental permeability. Large pulmonary conduit vessels and alveolar capillary segments have different intrinsic fluid and protein permeabilities and may experience differing degrees of injury due to differences in endothelial phenotypes (20, 64, 124). These segmental differences have been evaluated in rat lungs (4, 71) but have not been studied in mice. Most studies have used a sustained increase in airway pressure to collapse alveolar septal vessels (zone 1) and separately measure Kf for arterial (Kf,a) and venous (Kf,v) segments by raising their respective reservoirs (4). These are subtracted from the total Kf (Kf,t), which is assumed to equal Kf,c, to calculate that of the microvascular segment (Kf,mv) where
 | (5) |
Parker and Yoshikawa (124) recently used this method to partition Kf,c in isolated rat lungs into contributions from arterial (18%), microvascular (41%), and venous (41%) segments. Lung injury induced by hydrochloric acid, oleic acid, air emboli, ischemia-reperfusion, or high airway pressure each produced a different pattern of regional segmental injury indicating selective endothelial targeting (72, 85, 124). Although this approach yields a quantitative assessment of segmental permeability, it is important to recognize that Kf,a and Kf,v necessarily contain some component of filtration from very small conduit vessels and corner alveolar vessels. This is because the mechanical stresses at the alveolar corners permit some flow, even when alveolar pressure exceeds arterial pressure (Ppa) by 8-16 cmH2O (84). An alternative approach for measuring segmental vascular filtration in isolated rat lungs based on a stop-flow technique has been proposed by Lin et al. (91) in which the relative concentration increase for an impermeant dextran marker was used to assess filtration rate. With the use of this method, both the baseline Kf,c and filtration resulting from hydrochloric acid injury were equally distributed across venous and alveolar capillary segments, with only a minor contribution from arterial segments.
Direct measurements of Lp in small (40-µm) pulmonary venules and arterioles allow more specific evaluation of regional heterogeneity in hydraulic permeability. Bhattacharya (12) has applied the split-drop technique to measure Lp in subpleural surface pulmonary vessels. With the use of micropuncture, an oil droplet placed in a surface vessel is split with a droplet of buffer, and the rate of fluid absorption is determined by videomicroscopy. Assuming that the filtration S was approximated by the area of a cylinder, venular Lp averaged 2.9 ± 0.02 x 10-7 ml·s-1·cmH2O-1·cm-2 (12). Lp for surface venules exceeded that of surface arterioles by 40% in a comparative study using the split-drop technique (135).
Evaluation of pulmonary hemodynamics. Pulmonary vascular compliance (Cvas) and total vascular resistance (Rt) may also be altered by lung injury. Cvas will decrease and Rt will increase with vascular damage due to occlusive emboli and compression by alveolar edema as well as vasoconstriction. However, an increase in pulmonary vascular pressure will reduce both Cvas and Rt even in uninjured lungs (112, 141). Cvas can be estimated using the first 30 s of the lung weight gain curve by
 | (6) |
while Rt is calculated using the perfusate flow (Q), Ppa, and Ppv
 | (7) |
Typical Cvas values in isolated rat and mouse lung (
5.5 ml·min-1·100 g-1) (112, 197) were higher compared with those in rabbit and dog lung (
4.0 and 2.2 ml·min-1·100 g-1, respectively) (53, 141, 174). Typical Rt values (cmH2O-1·liters-1·min-1·100 g-1) for isolated rat (
5) and mouse (
7-10) lungs were significantly higher than values measured in intact rats (
3.5) and mice (
3.2) (54, 112, 162, 186, 197). Increased vascular resistance due to significant vascular derecruitment with injury may result in a falsely negative Kf,c response, i.e., the increase in permeability is offset by the decrease in S (158, 177).
Vascular permeability to proteins. Transvascular transport of labeled proteins, such as albumin, can be more accurately used to determined vascular permeability in isolated lungs than in intact animals because Clr for albumin can be measured under defined filtration conditions. This allows the investigator to calculate actual membrane permeability coefficients such as
and PS. Typical
and PS values derived from albumin Clr in rat and mouse lung are summarized in Table 2. Kern and Malik (68), and more recently Rippe and Taylor (143), measured
and PS using a double isotope technique in isolated lungs. Figure 4 gives a schematic representation of the experimental procedure. Isogravimetric lungs (Jv = 0) were perfused with 125I-labeled albumin at a known concentration for 3-8 min, followed by 131I-labeled albumin infused during a high filtration state (Jv > 0) for a similar time period. Residual perfusate was then washed out of the circulation before evaluating tissue accumulation of the radiolabeled tracer proteins. Clr for each albumin isotope was calculated by
 | (8) |
where A is the tissue radioactivity normalized to initial LW or dry weight. Because albumin clearance due to convection is minimal under isogravimetric conditions, diffusion predominates, and the isogravimetric albumin clearance yields an estimate of PS
 | (9) |
Conversely, the albumin clearance during the high filtration state, Clrfiltration, is primarily due to convective albumin transport with a minimal contribution by diffusion (143). Note that a correction for overlapping exposure periods is necessary for clearances measured by the method of Rippe and Taylor (143). Thus at high filtration states, Eq. 3 reduces to
 | (10) |
and
 | (11) |
Kreienberg et al. (79) and Waypa et al. (190) have modified this analysis by combining values for 3-min albumin Clr measured at different filtration rates. A least squares regression analysis was used to fit these data to Eq. 3, yielding estimates of PS (intercept) and
(slope) for each experimental group. Figure 5 shows the increasing albumin Clrfiltration as Jv increases with different degrees of oxidant-induced vascular injury. The dramatic increases in paracellular convective protein transport that occur when
decreases reflects the increased protein conductance relative to water. The increase in PS as
decreases for all except the H2O2-only group is compatible with increased diffusion through progressive increases in pore S (137).

View larger version (14K):
[in this window]
[in a new window]
|
Fig. 4. Isolated perfused rat lung weight gain curve showing the method used by Rippe and Taylor (143) to evaluate permeability-surface area product and reflection coefficient ( ) in rat lung. Clr (protein clearance)isogravimetric was calculated by subtracting the Clrfiltration (crosshatched bar 2) from the total 125I albumin accumulation in lung (hatched bar 1) during the experiment.
|
|

View larger version (24K):
[in this window]
[in a new window]
|
Fig. 5. Albumin clearances in isolated perfused rat lungs measured as a function of filtration rate. Solid lines indicate the least squares fit of Eq. 3 with the data in each treatment group. Note that albumin clearances increase as a function of both increased filtration rate and increased permeability (decreased ). Js, protein flux; Cp, plasma protein concentration; Jv, fluid filtration. [From Waypa et al. (190).]
|
|
Osmotic transients. When a hyperosmotic crystalloid solution is suddenly added to the lung perfusate to increase crystalloid osmotic pressure (
), water moves from the cells, interstitium, and alveolar space into the circulation through both transcellular and paracellular pathways (16, 164). A reflection coefficient can be calculated for the osmotic solute from Eq. 1 by assuming that the initial rate of weight loss, (
W/
t)0, is equal to the osmotically driven water flow, where
 | (12) |
Because crystalloid-induced water movement occurs primarily through the cells, the aquaporins (AQP) mediate much of the flow. These proteins have been identified in the lung microvessels (AQP-1), apical membrane of type I epithelium (AQP-5), and basolateral membrane of airway epithelium (AQP-4) (182). Carter et al. (16) introduced a novel method for evaluating alveolar fluid concentration and transcellular water permeability in isolated mouse lungs based on changes in surface fluorescence of an intra-alveolar FITC-dextran indicator. However, these methods have limited usefulness for evaluating lung injury because nearly all of the filtration in injured lungs occurs through paracellular pathways. In earlier studies using equivalent pore analysis of lymph data, we observed that only 3-5% of lung vascular filtration could be accounted for through water-only pathways during high filtration states, even in uninjured lungs (117). More recently, Song et al. (161) found no difference in edema accumulation in lungs of mice deficient in AQP-1, AQP-4, or AQP-5 after injury induced by HCl aspiration, thiourea infusion, or hyperoxia, and attributed these results to a predominance of Jv through paracellular pathways during injury.
In contrast, a reverse osmotic transient approach based on dilution of plasma proteins may be useful for evaluating injury in isolated perfused rat or mouse lungs because filtration across the capillaries is accompanied by exclusion of proteins from junctional paracellular pathways (144). From an isogravimetric state, the initial increase in filtration (
W/
t)0 is recorded at constant vascular pressures when the protein osmotic pressure of the perfusate is suddenly decreased by dilution with buffer (
P). The effective measured protein osmotic pressure across the lung capillaries (
M) is obtained using a previously determined Kf,c value where
 | (13) |
and thus
can be calculated using
 | (14) |
where 
P is the actual total protein oncotic pressure measured using an osmometer. Baseline
for total proteins in uninjured rat lungs was 0.75. In injured lungs where an isogravimetric baseline state cannot be obtained, (
W/
t)0 can be calculated as the difference between baseline and dilution-induced filtration rates.
Integral mass balance method. Maron and Pilati (94, 131) first showed that vascular permeability can be assessed using a solvent drag reflection coefficient for total protein (
f) in blood-perfused isolated lungs without the use of radioactive isotopes based on the relative transvascular movements of water and protein across the vascular barrier. This approach was subsequently modified and critically evaluated by Wolf et al. (194). With the use of the increase in hematocrit in the reservoir blood as a indicator of water filtration relative to the change in reservoir protein concentration during a period of sustained filtration,
f can be calculated using
 | (15) |
where H1 and C1 are hematocrit and total protein concentration, respectively, under baseline conditions, H2 and C2 are hematocrit and total protein concentration after a period of sustained filtration induced by increased vascular pressure, and C* is the average protein concentration over the filtration period. In pump-perfused lungs, corrections for errors in hematocrit and perfusate protein concentration due to red blood cell hemolysis are needed (96). This method has been applied to evaluate lung vascular permeability in isolated canine lung after vascular injury (93, 95, 123, 174, 176) but has not been used in mice. This method requires filtration of
10% of the perfusate volume to obtain changes in protein concentration and hematocrit that can be reliably measured.
Protein permeability in lungs of intact animals. Czermak et al. (24) and Lentsch et al. (87) have used a modified albumin clearance as a "permeability index" to study lung injury induced by immune complexes in intact rats. In this model, 10 µCi of 125I BSA are given with 10 mg of cold BSA after intratracheal instillation of 1.5 mg of anti-BSA antibodies, and then the permeability index is calculated using Eq. 8. In these studies, the permeability index averaged
0.2 for controls and 0.4 after immune complex injury at 4 h. Because of the long exposure times, the permeability index will not represent the initial albumin clearance but will reflect cumulative albumin leak into the lung interstitium relative to uninjured animals over the same time course. Normalization of lung radioactivity to initial lung weight and/or body weight is critical to compensate for the larger vascular S in larger animals. However, inbred strains generally have consistent LW/BW ratios over a given weight range (115), although the ratio can vary between strains. The permeability index will overestimate actual Clrisogravimetric if pulmonary vascular pressures and convective albumin flux increase. Similarly, endothelial permeability will be underestimated in the presence of vascular occlusion with derecruitment and return of labeled albumin to the circulation via pulmonary lymphatics (158). Although changes in the blood hematocrit or residual blood in lung tissue could affect variability, the permeability index has been a useful tool for detection of lung injury in rat models. Extravasation of radiolabeled albumin has also been used effectively to track lung permeability increases in selectin-deficient mice (192), during pancreatitis-induced lung injury in granulocyte/macrophage colony-stimulating factor-deficient mice (35), in septic mice expressing an adhesion molecule inhibitory factor (196), and after acid aspiration injury in C5-deficient mice (81).
Lung albumin extravasation has also been measured in mice using a dual isotope technique where 125I human serum albumin (HSA) was infused via a tail vein and 131I HSA was infused as a plasma volume marker immediately before death (102). This method was also used to assess airway albumin extravasation by counting radioactivity in excised airways (31) and increases in epithelial and endothelial permeability using the reverse flux of intratracheal radiolabeled albumin to the blood (89).
Evans blue dye has been proposed as an alternative to radioisotopes for measurement of albumin extravasation in airways and systemic organs since there was a high correlation between the Evans blue dye-labeled albumin and radiolabeled albumin methods (10, 44). Patterson et al. (126) reported excellent correlations of Evans blue dye-labeled albumin flux with that of 125I-labeled albumin and endogenous albumin when albumin flux was measured in lavage fluid from isolated rat lungs and across endothelial monolayers. However, in other studies, albumin clearances in rat lungs were approximately fourfold higher using the dye-albumin conjugate compared with 125I-labeled albumin even though unbound dye was not detected (25). A further limitation is that the absorbance of the dye must be corrected for residual heme pigments. Nonetheless, clearances of Evans blue-labeled albumin have been effectively used to detect increases in vascular permeability in rat models of lung injury (188).
Bronchoalveolar lavage. Bronchoalveolar lavage (BAL) has been used extensively to detect lung injury because plasma proteins, inflammatory cytokines, and inflammatory cells are present in very low or undetectable amounts in the alveolar spaces of normal lungs. Significant increases in BAL plasma proteins generally indicate a loss of endothelial barrier function. Altered concentrations of epithelial cell-specific protein markers can also be used as an index of epithelial damage (51). In patients with acute lung injury and ARDS, BAL surfactant proteins (SP) A, B, and D were decreased, whereas Clara cell secretory protein (CC16) generally increased, but responses varied with the disease present (51). In rats and mice, BAL CC16 was increased in virus infection and ozone exposure but decreased in most other types of lung injury (5, 6, 52, 88). Measurement of BAL lactate dehydrogenase has also been used to evaluate nonspecific injury to airway and alveolar epithelial cells (80).
Recent applications of proteomics promise to revolutionize the application of BAL analysis to disease diagnosis (107, 108). The several hundred proteins identified in BAL include proteins related to lipid metabolism, immune response, innate immunity, oxidative stress, proteases and antiproteases, cell proliferation, tissue repair, and intracellular proteins released by cell lysis (107, 108). Lung-specific cell markers of cell function may vary, depending on the type of injury, but some 45% of 460 BAL proteins were plasma derived and could be potential vascular injury markers (168). BAL cells from uninjured lungs consist of
85% alveolar macrophages, 10% lymphocytes, and smaller numbers of neutrophils and eosinophils. Rats and mice show marked increases in neutrophil numbers and complement components after LPS or acid challenge (40, 81, 192, 196).
Protein leak across the alveolar-capillary barrier can be assessed by measurement of total protein concentration in BALF using colorimetric or densitometric methods (66). BAL protein concentrations can be affected by low cardiac output, vessel occlusion, and high filtration pressures in addition to altered permeability. BAL albumin may be more sensitive to increases in permeability than total protein because albumin has a discrete molecular radius (3.7 nm) that is smaller than the average radius for total proteins (
5.5 nm). Albumin can be assessed using colorimetric assays, albumin labeled with radioisotopes, or Evans blue dye or a specific ELISA assay (81, 82, 113, 198). In addition, the intensity of an inflammatory response can be monitored by total and differential cell counts as well as cytokine and chemokine levels in BALF (49, 66, 181). In small animals, BAL is generally performed as a single measurement after death and has been used to evaluate lung injury in response to ozone in iNOS (66), Toll receptor, and TNF-
receptor-deficient mice (21, 75, 76) and in mice after sepsis (80). However, methods for serial BAL in rats and mice have recently been reported. Varner et al. (179) compared BAL inflammatory cell composition of left lower lobe with whole lung lavage in uninjured lungs of rats. Cell numbers and percentages were generally proportional between lung segments and whole lungs, but some significant differences in differential cell count percentages were observed. Segmental lavages could be repeated at weekly intervals and were well tolerated. Brown et al. (14) reported a method for endotracheal intubation of mice, which was used by Walters et al. (187) to perform serial whole lung BAL measurements on mice at daily and weekly intervals using 0.6 ml of Hanks' salt solution. The mice tolerated the serial BAL, but there were significant increases in airway resistance and decreases in lung compliance immediately after lavage. Small but significant differences in BAL total protein and cell composition occurred over time, and the response in the different mouse strains varied after the serial lavages. This is a promising method because each animal serves as its own control, minimizing intrastrain and individual variations.
Lung-specific protein markers in plasma. The appearance in plasma of secreted proteins specific to lung epithelium has been used to indicate breakdown of the air-blood barrier during lung injury (51). These proteins are normally present in very low levels in plasma and have at least three orders of magnitude higher concentrations in airway fluid in rats and mice, which facilitates rapid diffusion into blood in the presence of lung injury (45). Marker proteins include Clara cell secretory protein (CC10 and CC16,
16 kDa), SP-A (
28 kDa), SP-B (
40 kDa), SP-D (
40 kDa), and mucin-associated antigens (>200 kDa). In patients with acute lung injury, increases in serum levels have been reported threefold for CC16, twofold for SP-A, fivefold for SP-B, and fivefold for mucin antigens (51). Interpretation of plasma levels of lung-specific proteins can be complicated by changes in secretion rates induced by injury, changes in renal clearance of CC16, and aggregation of surfactant proteins to form multimers or combine with surfactant lipids (51). Doyle et al. (26) reported that plasma CC16 levels correlated better with renal function and SP-A and SP-B with histological indices of lung injury in patients with acute respiratory failure, although all three correlated with the decrease in blood oxygenation. In rats ventilated at sufficiently large tidal volumes to produce lung injury, blood CC16 levels correlated with tidal volume and BALF total protein concentration (88), whereas rats injured using intratracheal LPS also had blood CC16 levels that correlated with LPS dose as well as BALF total protein concentrations in both control and injury groups (6). In mice injured by high lung inflation pressures, plasma CC16 increased with minimal lung damage and was correlated with peak inflation pressure and lung wet/dry weight as well as BALF albumin and total protein concentrations (199). Other recently described proteins specific for lung epithelial cells may also prove to be useful markers of injury. These include the type I cell markers, rT140 in rats and HTI56 in humans (97, 106), the rat type II antigen for MMC4 (13), and the Clara cell secretoglobins SCGB3A1 and SCGB3A2 (139). Plasma rT140 and EVLW increased in proportion to peak inflation pressure, and microscopic evidence of injury in rats ventilated at high and low tidal volumes after acid aspiration (34). Thus small lung-specific proteins such as CC16 can increase early in lung injury but larger proteins, which are not cleared by the kidney, may prove more reliable for assessing the degree of injury.
Mechanisms of Protein Transport
There is overwhelming evidence that capillary filtration rate is closely coupled to capillary hydrostatic pressure and that protein transport is coupled to the filtration rate (99, 140, 166, 167). Transvascular protein clearance of 3.7- to 12.0-nm protein fractions in canine lung over a range of lymph flows was directly related to lymph flow and inversely related to protein molecular radius. These data best fit two protein sieving pores of 8.0 nm and 20.0 nm radius and a small number of water-only pores (117). In a kinetic analysis of labeled albumin equilibration in dog lung lymph based on Eq. 2, a baseline lung lymph flow of 0.067 ml·min-1·100 g-1 lung tissue was calculated with an albumin PS of 0.061 ml·min-1·100 g-1. Thus dissipative albumin transport would largely account for the simultaneously measured endogenous lung lymph albumin concentration at baseline filtration rates that equals 80% that of plasma (121). In an isogravimetric isolated lung, essentially all albumin clearance (Clrisogravimetric) would occur by such dissipative processes. However, application of Eq. 2 also predicts that diffusion or other dissipative processes would be maximal at a Pe of
0.6 for the
predicted for lung but would become minimal as lymph flow increased to a Pe of 3 (137). Dissipative albumin transport has been attributed to diffusion down a concentration gradient, vesicular transcytosis, or volume circulation (99, 120, 140, 166). Clearance by volume circulation would be driven by the oncotic pressure gradient (120), but experimental methods have not been devised to reliably discriminate among these three mechanisms.
Since the description of vesicles in endothelial cells more than 40 years ago, microscopists have postulated a role in transcapillary transport of plasma proteins from blood to tissue (101, 110, 160). The rapid uptake of labeled proteins by vesicles and the subsequent appearance of these proteins on the abluminal side of capillaries have been used to support the concept of transcytosis, or the active shuttle of proteins across endothelium by vesicles. Transient fusion of vesicles may form a transcellular pathway across the endothelial cell that would allow convective transport of fluid and protein and could serve as the large pore in capillaries (160).
We now know that caveolae and clathrin-coated vesicles possess all of the molecular machinery to transport proteins within the cell (136, 145). Caveolins and cholesterol are necessary for calveolae formation, and GTP binding proteins are necessary to recruit vesicle-coating proteins and supply energy for budding. Targeting soluble N-ethylmaleimide-attachment receptor (SNARE) proteins on the vesicles bind to proteins on specific target membranes. Specific receptors on the protein coat can sort and concentrate proteins, and specific peptide domains supply targeting signals for transport between Golgi apparatus, endoplasmic reticulum, endosomes, and apical and basal cell membranes (145). Schnitzer and Oh (150) showed high-affinity binding of albumin to albondin, a gp60 protein in endothelial cell vesicles, and proposed a major role for receptor-mediated albumin transport across lung endothelium. This 57- to 60-kDa protein was subsequently purified by Tiruppathi et al. (170). John et al. (58, 59) observed that endothelial monolayer albumin permeability could be reduced by depletion of gp60 or increased by cross-linking gp60. However, the reported binding constant of this protein would require a tissue albumin concentration of 10% of that in plasma for release of albumin to tissue (99), and the baseline lung lymph albumin concentration averages 80% of that in plasma (117). Several investigators have reported a reduced albumin uptake, and in some cases, a reduced transport across endothelial monolayers after inhibiting vesicle formation by treatment with N-ethylmaleimide (NEM) to bind fusion proteins or filipin, a cholesterol scavenger (134, 148, 150, 151, 185). In isolated rat lungs, Schnitzer et al. (151) and Vogel et al. (185) reported that vesicle inhibition with these agents reduced the PS of labeled albumin without affecting inulin uptake or Kf,c. However, isolated rat lung studies by Rippe and Taylor (143) could not confirm this effect and reported increases in isogravimetric albumin clearance (PS) and decreases in
at high doses of filipin and NEM in their preparation.
A detailed review of the vesicular transport controversy has recently been presented by Rippe et al. (142). Arguments against a significant contribution of active transcytosis include the strict first order kinetics of equilibration of tracer albumin between plasma and lung lymph because transcytosis should exhibit saturable Michaelis-Menten kinetics (111, 121), filtration-dependent clearance of both radiolabeled and endogenous albumin in lung lymph (56), the inability of cooling to block albumin clearance when active processes should cease (143), the inability of glutaraldehyde fixation to block albumin clearance (142), and the lack of any disturbance in the plasma-tissue distribution of plasma proteins, even across the blood-brain barrier, in caveolin-1 knockout mice that totally lack caveoli (27).
There is little doubt that convective paracellular protein exchange predominates at high filtration rates, which is especially true during lung injury. In
-naphthyl thiourea-injured dog lungs, Rutili et al. (146) observed a 10-fold increase in lung lymph protein clearance for the same increase in left atrial pressure. Estimated pore sizes increased to 9.5 and 28.0 nm with 45% of protein clearance through the large pore population. The contribution of dissipative processes to protein clearance with this massive protein leak would be too small to detect. Although transcytosis likely occurs, such protein transport probably relates to targeted transfer of specific proteins rather than control of the transcapillary oncotic pressure gradient involved in fluid homeostasis. Caveoli are essential for supplementing cell membrane during stretch and repairing small plasma membrane breaks (142, 183). These small "signalosomes" are involved in vascular nitric oxide regulation, cell proliferation, and numerous signaling processes whose functions are yet to be discovered. Although vesicles can shuttle specific proteins to specified intracellular domains for processing (145), transcytosis does not appear to contribute significantly to the transcapillary protein leak that occurs during lung vascular injury.
Transcytosis has also been proposed as a mechanism involved in protein transport across the airway and alveolar epithelium in the lung (73). Estimates of alveolar epithelial equivalent pore radii of 0.5-1.0 nm (165) with very infrequent larger pores (169) could account for the normally low concentration of plasma proteins in alveolar lining fluid. Although asymmetric and protein-specific uptake studies suggest protein-specific binding proteins (73), the size- and concentration-dependent clearance of solutes from alveolar fluid suggests water-filled paracellular pores as the major route for solute clearance (29, 51). Thus the lung-to-blood transport of lung-specific proteins during injury likely occurs by diffusion down a concentration gradient through paracellular aqueous pathways where the severity of injury determines the potential pore area for diffusive transfer.
 |
SUMMARY
|
---|
Lung injury is a broad descriptor that can be applied to conditions ranging from mild interstitial edema without cellular injury to massive and fatal destruction of the lung. The small size of rats and mice places limits on the methods that can be used practically to assess injury. Accumulation of pulmonary edema can be easily and quantitatively measured using gravimetric methods. Corrections for residual blood volumes and excess Js or hemorrhage further improve the accuracy. Imaging methods are generally qualitative and do not easily resolve fine structure in mouse lung. Lung compliance measurements are repeatable and well correlated with edema but may be influenced by other factors, such as atelectasis and surfactant inactivation. Light and EM can be used regardless of lung size to detect edema as well as the nature and distribution of structural damage but are less useful for evaluating global lung dysfunction. Although increases in microvascular permeability in lung are always indicative of a pathological process, increases in fluid and/or Js due to increased permeability must be separated from those due to increased filtration pressure. An increase in the initial albumin clearance compared with controls matched for size and filtration pressure is a reliable indicator of vascular dysfunction, but alterations in the membrane coefficients Kf,c,
, and PS are the most accurate indicators of permeability. Generally, PS and Kf,c will increase and
will decrease with vascular injury, but combinations of different-sized paracellular pathways may lead to proportional differences in these changes. In addition, derecruitment of S may affect PS and Kf,c but would not alter measurements of
.
 |
ACKNOWLEDGMENTS
|
---|
The authors appreciate the constructive comments of Dr. Aubrey E. Taylor.
GRANTS
This work was supported by National Heart, Lung, and Blood Institute Grants HL-66299 and HL-61955.
 |
FOOTNOTES
|
---|
Address for reprint requests and other correspondence: J. C. Parker, Dept. of Physiology, MSB 3074, Univ. of South Alabama, Mobile, AL 36688-0002 (E-mail: jparker{at}usouthal.edu).
 |
REFERENCES
|
---|
- Aberle DR, Wiener-Kronish JP, Webb WR, and Matthay MA. Hydrostatic versus increased permeability pulmonary edema: diagnosis based on radiographic criteria in critically ill patients. Radiology 168: 73-79, 1988.[Abstract]
- Adams AB, Cakar N, and Marini JJ. Static and dynamic pressure-volume curves reflect different aspects of respiratory system mechanics in experimental acute respiratory distress syndrome. Respir Care 46: 686-693, 2001.[Medline]
- Adkins WK, Barnard JW, May MJ, Seibert AF, Haynes J, and Taylor AE. Compounds that increase cAMP prevent ischemia reperfusion pulmonary capillary injury. J Appl Physiol 72: 492-497, 1992.[Abstract/Free Full Text]
- Albert RK, Kirk W, Pitts C, and Butler J. Extra-alveolar vessel fluid filtration coefficients in excised and in situ canine lobes. J Appl Physiol 59: 1555-1559, 1985.[Abstract/Free Full Text]
- Arsalane K, Broeckaert F, Knoops B, Clippe A, Buchet JP, and Bernard A. Increased serum and urinary concentrations of lung clara cell protein in rats acutely exposed to ozone. Toxicol Appl Pharmacol 159: 169-174, 1999.[CrossRef][ISI][Medline]
- Arsalane K, Broeckaert F, Knoops B, Wiedig M, Toubeau G, and Bernard A. Clara cell specific protein (CC16) expression after acute lung inflammation induced by intratracheal lipopolysaccharide administration. Am J Respir Crit Care Med 161: 1624, 2000.[Abstract/Free Full Text]
- Aukland K and Reed RK. Interstitial-lymphatic mechanisms in the control of extracellular fluid volume. Physiol Rev 73: 1-78, 1993.[Abstract/Free Full Text]
- Bachofen H, Schurch S, and Weibel ER. Experimental hydrostatic pulmonary edema in rabbit lungs. Barrier lesions. Am Rev Respir Dis 147: 997-1004, 1993.[ISI][Medline]
- Barazzone C, Donati YR, Rochat AF, Vesin C, Kan CD, Pache JC, and Piguet PF. Keratinocyte growth factor protects alveolar epithelium and endothelium from oxygen-induced injury in mice. Am J Pathol 154: 1479-1487, 1999.[Abstract/Free Full Text]
- Barnes PJ, Boschetto P, and Barnes PJ. Plasma exudation: correlation between Evans Blue dye and radiolabeled albumin in guinea pig airways in vivo. J Pharmacol Methods 21: 309-315, 1989.[CrossRef][ISI][Medline]
- Beckmann N, Tigani B, Mazzoni L, and Fozard JR. MRI of lung parenchyma in rats and mice using a gradient-echo sequence. NMR Biomed 14: 297-306, 2001.[CrossRef][ISI][Medline]
- Bhattacharya J. Hydraulic conductivity of lung venules determined by split-drop technique. J Appl Physiol 64: 2562-2567, 1988.[Abstract/Free Full Text]
- Boylan GM, Pryde JG, Dobbs LG, and McElroy MC. Identification of a novel antigen on the apical surface of rat alveolar epithelial type II and Clara cells. Am J Physiol Lung Cell Mol Physiol 280: L1318-L1326, 2001.[Abstract/Free Full Text]
- Brown RH, Walters DM, Greenberg RS, and Mitzner W. A method of endotracheal intubation and pulmonary functional assessment for repeated studies in mice. J Appl Physiol 87: 2362-2365, 1999.[Abstract/Free Full Text]
- Brown VM, Ossadtchi A, Khan AH, Gambhir SS, Cherry SR, Leahy RM, and Smith DJ. Gene expression tomography. Physiol Genomics 8: 159-167, 2002.[Abstract/Free Full Text]
- Carter EP, Matthay MA, Farinas J, and Verkman AS. Transalveolar osmotic and diffusional water permeability in intact mouse lung measured by a novel surface fluorescence method. J Gen Physiol 108: 133-142, 1996.[Abstract]
- Caruthers SD, Paschal CB, Pou NA, Roselli RJ, and Harris TR. Regional measurements of pulmonary edema by using magnetic resonance imaging. J Appl Physiol 84: 2143-2153, 1998.[Abstract/Free Full Text]
- Champion HC, Villnave DJ, Tower A, Kadowitz PJ, and Hyman AL. A novel right-heart catheterization technique for in vivo measurement of vascular responses in lungs of intact mice. Am J Physiol Heart Circ Physiol 278: H8-H15, 2000.[Abstract/Free Full Text]
- Cherry SR and Gambhir SS. Use of positron emission tomography in animal research. ILAR J 42: 219-232, 2001.[Medline]
- Chetham PM, Babal P, Bridges JP, Moore TM, and Stevens T. Segmental regulation of pulmonary vascular permeability by store-operated Ca2+ entry. Am J Physiol Lung Cell Mol Physiol 276: L41-L50, 1999.[Abstract/Free Full Text]
- Cho HY, Zhang LY, and Kleeberger SR. Ozone-induced lung inflammation and hyperreactivity are mediated via tumor necrosis factor-
receptors. Am J Physiol Lung Cell Mol Physiol 280: L537-L546, 2001.[Abstract/Free Full Text]
- Christofidou-Solomidou M, Scherpereel A, Solomides CC, Christie JD, Stossel TP, Goelz S, and DiNubile MJ. Recombinant plasma gelsolin diminishes the acute inflammatory response to hyperoxia in mice. J Investig Med 50: 54-60, 2002.[ISI][Medline]
- Costello ML, Mathieu-Costello O, and West JB. Stress failure of alveolar epithelial cells studied by scanning electron microscopy. Am Rev Respir Dis 145: 1446-1455, 1992.[ISI][Medline]
- Czermak BJ, Breckwoldt M, Ravage ZB, Huber-Lang M, Schmal H, Bless NM, Friedl HP, and Ward PA. Mechanisms of enhanced lung injury during sepsis. Am J Pathol 154: 1057-1065, 1999.[Abstract/Free Full Text]
- Dallal MM and Chang SW. Evans blue dye in the assessment of permeability-surface are product in perfused rat lungs. J Appl Physiol 77: 1030-1035, 1994.[Abstract/Free Full Text]
- Doyle IR, Hermans C, Bernard A, Nicholas TE, and Bersten AD. Clearance of Clara cell secretory protein 16 (CC16) and surfactant proteins A and B from blood in acute respiratory failure. Am J Respir Crit Care Med 158: 1528, 1998.[Abstract/Free Full Text]
- Drab M, Verkade P, Elger M, Kasper M, Lohn M, Lauterbach B, Menne J, Lindschau C, Mende F, Luft FC, Schedl A, Haller H, and Kurzchalia TV. Loss of caveolae, vascular dysfunction, and pulmonary defects in caveolin-1 gene-disrupted mice. Science 293: 2449-2452, 1928.[CrossRef]
- Drake RE and Taylor AE. Estimation of the filtration coefficient of pulmonary exchange vessels. Am J Physiol Heart Circ Physiol 234: H266-H274, 1978.[Abstract/Free Full Text]
- Egan EA, Nelson RM, and Oliver RE. Lung inflation and alveolar permeability to nonelectrolytes in the adult sheep in vivo. J Physiol 260: 409-424, 1976.[Abstract]
- Elliott AR, Fu Z, Tsukimoto K, Prediletto R, Mathieu-Costello O, and West JB. Short-term reversibility of ultrastructural changes in pulmonary capillaries caused by stress failure. J Appl Physiol 73: 1150-1158, 1992.[Abstract/Free Full Text]
- Erjefalt JS, Andersson P, Gustoafsson B, Korsgren M, Sonmark B, and Persson CGA. Allergen challenge induced extravasation of plasma in mouse airways. Clin Exp Allergy 28: 1013-1020, 1998.[CrossRef][ISI][Medline]
- Ewart S, Levitt R, and Mitzner W. Respiratory system mechanics in mice measured by end-inflation occlusion. J Appl Physiol 79: 560-566, 1995.[Abstract/Free Full Text]
- Fang X, Fukuda N, Barbry P, Sartori C, Verkman AS, and Matthay MA. Novel role for CFTR in fluid absorption from the distal air spaces of the lung. J Gen Physiol 119: 199-207, 2002.[Abstract/Free Full Text]
- Frank JA, Gutierrez JA, Jones KD, Allen L, Dobbs L, and Matthay MA. Low tidal volume reduces epithelial and endothelial injury in acid-injured rat lungs. Am J Respir Crit Care Med 165: 242-249, 2002.[Abstract/Free Full Text]
- Frossard JL, Saluja AK, Mach N, Lee HS, Bhagat L, Hadenque A, Rubbia-Brandt L, Dranoff G, and Steer ML. In vivo evidence for the role of GM-CSF as a mediator in acute pancreatitis-associated lung injury. Am J Physiol Lung Cell Mol Physiol 283: L541-L548, 2002.[Abstract/Free Full Text]
- Fujimoto K, Parker JC, and Kayes SG. Activated eosinophils increase vascular permeability and resistance in isolated perfused rat lungs. Am Rev Respir Dis 142: 1414-1421, 1990.[ISI][Medline]
- Fukuda N, Folkesson HG, and Matthay MA. Relationship of interstitial fluid volume to alveolar fluid clearance in mice: ventilated vs. in situ studies. J Appl Physiol 89: 672-679, 2000.[Abstract/Free Full Text]
- Gannon BJ. Preparation of microvascular corrosion casting media: procedure for partial polymerization of methyl methacrylate using ultra-violet light. Biomed Res 2: 227-233, 1981.[ISI]
- Gao X, Kouklis P, Xu N, Minshall RD, Sandoval R, Vogel SM, and Malik AB. Reversibility of increased microvessel permeability in response to VE-cadherin disassembly. Am J Physiol Lung Cell Mol Physiol 279: L1218-L1225, 2000.[Abstract/Free Full Text]
- Gao X, Xu N, Sekosan M, Mehta D, Ma SY, Rahman A, and Malik AB. Differential role of CD18 integrins in mediating lung neutrophil sequestration and increased microvascular permeability induced by Escherichia coli in mice. J Immunol 167: 2895-2901, 2001.[Abstract/Free Full Text]
- Geiser M, Zimmermann B, Baumann M, and Cruz-Orive LM. Does lack of Cftr gene lead to developmental abnormalities in the lung? Exp Lung Res 26: 551-564, 2000.[CrossRef][ISI][Medline]
- Glaab T, Daser A, Braun A, Neuhaus-Steinmetz U, Fabel H, Alarie Y, and Renz H. Tidal midexpiratory flow as a measure of airway hyperresponsiveness in allergic mice. Am J Physiol Lung Cell Mol Physiol 280: L565-L573, 2001.[Abstract/Free Full Text]
- Granger DN and Taylor AE. Permeability of intestinal capillaries to endogenous macromolecules. Am J Physiol Heart Circ Physiol 238: H457-H464, 1980.[Abstract/Free Full Text]
- Green TP, Johnson DE, Marchessault RP, and Gatto CW. Transvascular flux and tissue accrual of Evans blue: effects of endotoxin and histamine. J Lab Clin Med 111: 173-183, 1988.[ISI][Medline]
- Halatek T, Hermans C, Broeckaert F, Wattiez R, Wiedig M, Toubeau G, Falmagne P, and Bernard A. Quantification of Clara cell protein in rat and mouse biological fluids using a sensitive immunoassay. Eur Respir J 11: 726-733, 1998.[Abstract/Free Full Text]
- Hamelmann E, Schwarze J, Takeda K, Oshiba A, Larsen GL, and Irvin CG. Noninvasive measurement of airway responsiveness in allergic mice using barometric plethysmography. Am J Respir Crit Care Med 156: 766-775, 1997.[Abstract/Free Full Text]
- Hancock BJ, Landolfo KP, and Oppenheimer L. Slow phase of transvascular fluid flux reviewed. J Appl Physiol 69: 456-464, 1990.[Abstract/Free Full Text]
- Haraldsson B and Rippe B. Orosomucoid as one of the serum components contributing to normal capillary permselectivity in rat skeletal muscle. Acta Physiol Scand 129: 127-135, 1987.[ISI][Medline]
- Hardie WD, Prows DR, Leikauf GD, and Korfhagen TR. Attenuation of acute lung injury in transgenic mice expressing human transforming growth factor-
. Am J Physiol Lung Cell Mol Physiol 277: L1045-L1050, 1999.[Abstract/Free Full Text]
- Harris NR, Parker RE, Pou NA, and Roselli RJ. Canine pulmonary filtration coefficient calculated from optical, radioisotope, and weight measurements. J Appl Physiol 73: 2648-2661, 1992.[Abstract/Free Full Text]
- Hermans C and Bernard A. Lung epithelium-specific proteins. Characteristics and potential applications as markers. Am J Respir Crit Care Med 159: 646-678, 1999.[Free Full Text]
- Hermans C, Knoops B, Wiedig M, Arsalane K, Toubeau G, Falmagne P, and Bernard A. Clara cell protein as a marker of Clara cell damage and bronchoalveolar blood barrier permeability. Eur Respir J 13: 1014-1021, 1999.[Abstract/Free Full Text]
- Hernandez LA, Peevy KJ, Moise AA, and Parker JC. Chest wall restriction limits high airway pressure-induced lung injury in young rabbits. J Appl Physiol 66: 2364-2368, 1989.[Abstract/Free Full Text]
- Hill NS, Warburton RR, Pietras L, and Klinger JR. Nonspecific endothelin-receptor antagonist blunts monocrotaline-induced pulmonary hypertension in rats. J Appl Physiol 83: 1209-1215, 1997.[Abstract/Free Full Text]
- Hu X, Adamson RH, Liu B, Curry FE, and Weinbaum S. Starling forces that oppose filtration after tissue oncotic pressure is increased. Am J Physiol Heart Circ Physiol 279: H1724-H1736, 2000.[Abstract/Free Full Text]
- Ishibashi M, Reed RK, Townsley MI, Parker JC, and Taylor AE. Albumin transport across pulmonary capillary-interstitial barrier in anesthetized dogs. J Appl Physiol 70: 2104-2110, 1991.[Abstract/Free Full Text]
- Izbicki G, Segel MJ, Christensen TG, Conner MW, and Breuer R. Time course of bleomycin-induced lung fibrosis. Int J Exp Pathol 83: 111-119, 2002.[CrossRef][ISI][Medline]
- John TA, Vogel SM, Minshall RD, Ridge K, Tiruppathi C, and Malik AB. Evidence for the role of alveolar epithelial gp60 in active transalveolar albumin transport in the rat lung. J Physiol 533: 547-559, 2001.[Abstract/Free Full Text]
- John TA, Vogel SM, Tiruppathi C, Malik AB, and Minshall RD. Quantitative analysis of albumin uptake and transport in the rat microvessel endothelial monolayer. Am J Physiol Lung Cell Mol Physiol 284: L187-L196, 2003.[Abstract/Free Full Text]
- Johnstrom P, Harris NG, Fryer TD, Barret O, Clark JC, Pickard JD, and Davenport AP. 18F-endothelin-1, a positron emission tomography radioligand for the endothelin receptor system: radiosynthesis and in vivo imaging using microPET. Clin Sci 103, Suppl 48: 4S-8S, 2002.[ISI][Medline]
- Jones JE, Mendes L, Rudd MA, Russo G, Loscalzo J, and Zhang YY. Serial noninvasive assessment of progressive pulmonary hypertension in a rat model. Am J Physiol Heart Circ Physiol 283: H364-H371, 2002.[Abstract/Free Full Text]
- Kaner RJ, Ladetto JV, Singh R, Fukuda N, Matthay MA, and Crystal RG. Lung overexpression of the vascular endothelial growth factor gene induces pulmonary edema. Am J Respir Cell Mol Biol 22: 657-664, 2000.[Abstract/Free Full Text]
- Karau KL, Molthen RC, Dhyani A, Haworth ST, Hanger CC, Roerig DL, Johnson RH, and Dawson CA. Pulmonary arterial morphometry from microfocal X-ray computed tomography. Am J Physiol Heart Circ Physiol 281: H2747-H2756, 2001.[Abstract/Free Full Text]
- Kelly JJ, Moore TM, Babal P, Diwan AH, Stevens T, and Thompson WJ. Pulmonary microvascular and macrovascular endothelial cells: differential regulation of Ca2+ and permeability. Am J Physiol Lung Cell Mol Physiol 274: L810-L819, 1998.[Abstract/Free Full Text]
- Kennel SJ, Davis IA, Branning J, Pan H, Kabalka GW, and Paulus MJ. High resolution computed tomography and MRI for monitoring lung tumor growth in mice undergoing radioimmunotherapy: correlation with histology. Med Phys 27: 1101-1107, 2000.[CrossRef][ISI][Medline]
- Kenyon NJ, van der Vliet A, Schock BC, Okamoto T, McGrew GM, and Last JA. Susceptibility to ozone-induced acute lung injury in iNOS-deficient mice. Am J Physiol Lung Cell Mol Physiol 282: L540-L545, 2002.[Abstract/Free Full Text]
- Kern DF, Levitt D, and Wangensteen D. Endothelial albumin permeability measured with a new technique in perfused rabbit lung. Am J Physiol Heart Circ Physiol 245: H229-H236, 1983.[Abstract/Free Full Text]
- Kern DF and Malik AB. Microvascular albumin permeability in isolated perfused lung: effects of EDTA. J Appl Physiol 58: 372-375, 1985.[Abstract/Free Full Text]
- Khimenko PL, Bagby GJ, Fuseler J, and Taylor AE. Tumor necrosis factor-
in ischemic and reperfusion injury in rat lungs. J Appl Physiol 85: 2005-2011, 1998.[Abstract/Free Full Text]
- Khimenko PL, Barnard JW, Moore TM, Wilson PS, Ballard ST, and Taylor AE. Vascular permeability and epithelial transport effects on lung edema formation in ischemia and reperfusion. J Appl Physiol 77: 1116-1121, 1994.[Abstract/Free Full Text]
- Khimenko PL, Moore TM, Wilson PS, and Taylor AE. Role of calmodulin and myosin light chain kinase in lung ischemia-reperfusion injury. Am J Physiol Lung Cell Mol Physiol 271: L121-L125, 1996.[Abstract/Free Full Text]
- Khimenko PL and Taylor AE. Segmental microvascular permeability in ischemia reperfusion injury in rat lung. Am J Physiol Lung Cell Mol Physiol 276: L958-L960, 1999.[Abstract/Free Full Text]
- Kim KJ and Malik AB. Protein transport across the lung epithelial barrier. Am J Physiol Lung Cell Mol Physiol 284: L247-L259, 2003.[Abstract/Free Full Text]
- Kitamura Y, Hashimoto S, Mizuta N, Kobayashi A, Kooguchi K, Fujiwara I, and Nakajima H. Fas/FasL-dependent apoptosis of alveolar cells after lipopolysaccharide-induced lung injury in mice. Am J Respir Crit Care Med 163: 762-769, 2001.[Abstract/Free Full Text]
- Kleeberger SR, Reddy KK, Zhang LY, and Jedlicka AE. Genetic susceptibility to ozone induced lung hyperpermeability: role of toll-like receptor 4. Am J Respir Cell Mol Biol 22: 620-627, 2000.[Abstract/Free Full Text]
- Kleeberger SR, Reddy SP, Zhang LY, Cho HY, and Jedlicka AE. Toll-like receptor 4 mediates ozone-induced murine lung hyperpermeability via inducible nitric oxide synthase. Am J Physiol Lung Cell Mol Physiol 280: L326-L333, 2001.[Abstract/Free Full Text]
- Kobayashi H, Hataishi R, Mitsufuji H, Tanaka M, Jacobson M, Tomita T, Zapol WM, and Jones RC. Antiinflammatory properties of inducible nitric oxide synthase in acute hyperoxic lung injury. Am J Respir Cell Mol Biol 24: 390-397, 2001.[Abstract/Free Full Text]
- Kolb M, Bonniaud P, Galt T, Sime PJ, Kelly MM, Margetts PJ, and Gauldie J. Differences in the fibrogenic response after transfer of active transforming growth factor-
1 gene to lungs of "fibrosis-prone" and "fibrosis-resistant" mouse strains. Am J Respir Cell Mol Biol 27: 141-150, 2002.[Abstract/Free Full Text]
- Kreienberg PB, Vincent PA, Bell DR, Saba TM, and Minnear FL. Isoproterenol decreases protein permeability in edematous isolated rabbit lungs: estimation of PS and
. J Appl Physiol 77: 325-331, 1994.[Abstract/Free Full Text]
- Kristof AS, Goldberg P, Laubach V, and Hussain SN. Role of inducible nitric oxide synthase in endotoxin-induced acute lung injury. Am J Respir Crit Care Med 158: 1883-1889, 1998.[Abstract/Free Full Text]
- Kyriakides C, Austen W Jr, Wang Y, Favuzza J, Kobzik L, Moore FD Jr, and Hechtman HB. Membrane attack complex of complement and neutrophils mediate the injury of acid aspiration. J Appl Physiol 87: 2357-2361, 1999.[Abstract/Free Full Text]
- Kyriakides C, Favuzza J, Wang Y, Austen WG, Moore FD, and Hechtman HB. Recombinant soluble P-selectin glycoprotein ligand 1 moderates local and remote injuries following experimental lower torso ischaemia. Br J Surg 88: 825-830, 2001.[CrossRef][ISI][Medline]
- Lai YL and Chou HC. Respiratory mechanics and maximal expiratory flow in the anesthetized mouse. J Appl Physiol 88: 939-943, 2000.[Abstract/Free Full Text]
- Lamm WJ, Kirk KR, Hanson WL, Wagner WW Jr, and Albert RK. Flow through zone 1 lungs utilizes alveolar corner vessels. J Appl Physiol 70: 1518-1523, 1991.[Abstract/Free Full Text]
- Lamm WJ, Luchtel D, and Albert RK. Sites of leakage in three models of acute lung injury. J Appl Physiol 64: 1079-1083, 1988.[Abstract/Free Full Text]
- Lehnert S, Schreiner LJ, and el Khatib E. Factors influencing lung density in experimental models: results of studies using CT densitometry. Physiol Meas 14: 183-193, 1993.[CrossRef][ISI][Medline]
- Lentsch AB, Jordan JA, Czermak BJ, Diehl KM, Younkin EM, Sarma V, and Ward PA. Inhibition of NF-
B activation and augmentation of I
B
by secretory leukocyte protease inhibitor during lung inflammation. Am J Pathol 154: 239-247, 1999.[Abstract/Free Full Text]
- Lesur O, Hermans C, Chalifour JF, Picotte J, Levy B, Bernard A, and Lane D. Mechanical ventilation-induced pneumoprotein CC-16 vascular transfer in rats: effect of KGF pretreatment. Am J Physiol Lung Cell Mol Physiol 284: L410-L419, 2003.[Abstract/Free Full Text]
- Li XY, Donaldson K, and MacNee W. Lipopolysaccharide-induced alveolar epithelial permeability: the role of nitric oxide. Am J Respir Crit Care Med 157: 1027-1033, 1998.[Abstract/Free Full Text]
- Liao Y, Ishikura F, Beppu S, Asakura M, Takashima S, Asanuma H, Sanada S, Kim J, Ogita H, Kuzuya T, Node K, Kitakaze M, and Hori M. Echocardiographic assessment of LV hypertrophy and function in aortic-banded mice: necropsy validation. Am J Physiol Heart Circ Physiol 282: H1703-H1708, 2002.[Abstract/Free Full Text]
- Lin W, Jacobs E, Schapira RM, Presberg K, and Effros RM. Stop-flow studies of distribution of filtration in rat lungs. J Appl Physiol 84: 47-52, 1998.[Abstract/Free Full Text]
- Lundblad LK, Irvin CG, Adler A, and Bates JH. A reevaluation of the validity of unrestrained plethysmography in mice. J Appl Physiol 93: 1198-1207, 2002.[Abstract/Free Full Text]
- Maron MB. Effect of elevated vascular pressure transients on protein permeability in the lung. J Appl Physiol 67: 305-310, 1989.[Abstract/Free Full Text]
- Maron MB and Pilati CF. Calculation of the reflection coefficient from measurements of endogenous vascular indicators. J Appl Physiol 64: 1746-1748, 1988.[Abstract/Free Full Text]
- Maron MB and Pilati CF. Effect of papaverine on pulmonary vascular permeability to proteins. J Appl Physiol 65: 1367-1371, 1988.[Abstract/Free Full Text]
- Maron MB, Pilati CF, and Maender KC. Effect of hemolysis on reflection coefficient determined by endogenous blood indicators. J Appl Physiol 62: 2252-2257, 1987.[Abstract/Free Full Text]
- McElroy MC, Wiener-Kronish JP, Miyazaki H, Sawa T, Modelska K, Dobbs LG, and Pittet JF. Nitric oxide attenuates lung endothelial injury caused by sublethal hyperoxia in rats. Am J Physiol Lung Cell Mol Physiol 272: L631-L638, 1997.[Abstract/Free Full Text]
- Mercer RR, Russell ML, and Crapo JD. Alveolar septal structure in different species. J Appl Physiol 77: 1060-1066, 1994.[Abstract/Free Full Text]
- Michel CC and Curry FE. Microvascular permeability. Physiol Rev 79: 703-761, 1999.[Abstract/Free Full Text]
- Milne EN, Pistolesi M, Miniati M, and Giuntini C. The radiologic distinction of cardiogenic and noncardiogenic edema. AJR Am J Roentgenol 144: 879-894, 1985.[ISI][Medline]
- Minshall RD, Tiruppathi C, Vogel SM, and Malik AB. Vesicle formation and trafficking in endothelial cells and regulation of endothelial barrier function. Histochem Cell Biol 117: 105-112, 2002.[CrossRef][ISI][Medline]
- Miotla JM, Williams TJ, Hellewell PG, and Jeffery PK. A role for the
2 integrin CD11b in mediating experimental lung injury in mice. Am J Respir Cell Mol Biol 14: 363-373, 1996.[Abstract]
- Mitzner W, Brown R, and Lee W. In vivo measurement of lung volumes in mice. Physiol Genomics 4: 215-221, 2001.[Abstract/Free Full Text]
- Mitzner W and Tankersley C. Noninvasive measurement of airway responsiveness in allergic mice using barometric plethysmography. Am J Respir Crit Care Med 158: 340-341, 1998.[Medline]
- Moore TM, Shirah WB, Khimenko PL, Paisley P, Lausch RN, and Taylor AE. Involvement of CD40-CD40L signaling in postischemic lung injury. Am J Physiol Lung Cell Mol Physiol 283: L1255-L1262, 2002.[Abstract/Free Full Text]
- Newman V, Gonzalez RF, Matthay MA, and Dobbs LG. A novel alveolar type I cell-specific biochemical marker of human acute lung injury. Am J Respir Crit Care Med 161: 990-995, 2000.[Abstract/Free Full Text]
- Noel-Georis I, Bernard A, Falmagne P, and Wattiez R. Proteomics as the tool to search for lung disease markers in bronchoalveolar lavage. Dis Markers 17: 271-284, 2001.[ISI][Medline]
- Noel-Georis I, Bernard A, Falmagne P, and Wattiez R. Database of bronchoalveolar lavage fluid proteins. J Chromatog B Analyt Technol Biomed Life Sci 771: 221-236, 2002.[CrossRef][ISI]
- Omlor G, Niehaus GD, and Maron MB. Effect of peak inspiratory pressure on the filtration coefficient in the isolated perfused rat lung. J Appl Physiol 74: 3068-3072, 1993.[Abstract]
- Palade GE. Transport of quanta across the endothelium of blood capillaries (Abstract). Anat Rec 136: A254, 1960.
- Parker JC, Crain M, Grimbert F, Rutili G, and Taylor AE. Total lung lymph flow and fluid compartmentation in edematous dog lungs. J Appl Physiol 51: 1268-1277, 1981.[Abstract/Free Full Text]
- Parker JC, Drake DA, King JA, and Yoshikawa S. Ventilator induced lung injuryLeak sites identified using methacrylate vascular casts (Abstract). Am Rev Respir Crit Care Med 167: A775, 2003.
- Parker JC, Gillespie MN, Taylor AE, and Martin SL. Capillary filtration coefficient, vascular resistance and compliance in isolated mouse lungs. J Appl Physiol 87: 1421-1427, 1999.[Abstract/Free Full Text]
- Parker JC, Hamm CR, Coker-Flowers P, Strober M, and Parker JC. The neutralization of alveolar macrophages reduces barotrauma induced lung injury. FASEB J 16: A410-A410, 2002.
- Parker JC and Ivey CL. Isoproterenol attenuates high vascular pressure induced permeability increases in isolated rat lungs. J Appl Physiol 83: 1962-1967, 1997.[Abstract/Free Full Text]
- Parker JC, Ivey CL, and Tucker A. Gadolinium prevents high airway pressure induced permeability increases in isolated rat lungs. J Appl Physiol 84: 1113-1118, 1998.[Abstract/Free Full Text]
- Parker JC, Ivey CL, and Tucker A. Phosphotyrosine phosphatase and tyrosine inhibition modulate airway pressure induced lung injury. J Appl Physiol 85: 1753-1761, 1998.[Abstract/Free Full Text]
- Parker JC, Parker RE, Granger DN, and Taylor AE. Vascular permeability and transvascular fluid and protein transport in the dog lung. Circ Res 48: 549-561, 1981.[Abstract]
- Parker JC, Perry MA, and Taylor AE. Permeability of the microvascular barrier. In: Edema, edited by Staub NC and Taylor AE. New York: Raven, 1984, p. 143-187.
- Parker JC, Prasad R, Allison RA, Wojchiechowski WV, and Martin SL. Capillary filtration coefficients using laser densitometry and gravimetry in isolated dog lungs. J Appl Physiol 74: 1981-1987, 1993.[Abstract]
- Parker JC, Rippe B, and Taylor AE. Fluid filtration and protein clearances through large and small pore populations in dog lung capillaries. Microvasc Res 31: 1-17, 1986.[ISI][Medline]
- Parker JC, Ryan J, and Taylor AE. Plasma-lymph albumin kinetics, total lymph flow, and tissue hematocrit in normally hydrated dog lungs. Microvasc Res 28: 254-269, 1984.[ISI][Medline]
- Parker JC, Townsley MI, and Cartledge JT. Lung edema increases transvascular filtration rate but not filtration coefficient. J Appl Physiol 66: 1553-1560, 1989.[Abstract/Free Full Text]
- Parker JC, Townsley MI, Rippe B, Taylor AE, and Thigpen J. Increased microvascular permeability in dog lungs due to high peak airway pressures. J Appl Physiol 57: 1809-1816, 1984.[Abstract/Free Full Text]
- Parker JC and Yoshikawa S. Vascular segmental permeabilities at high peak inflation pressure in isolated rat lungs. Am J Physiol Lung Cell Mol Physiol 283: L1203-L1209, 2002.[Abstract/Free Full Text]
- Patlak CS, Goldstein DA, and Hoffman JF. The flow of solute and solvent across a two-membrane system. J Theor Biol 5: 426-442, 1963.[ISI][Medline]
- Patterson CE, Rhoades RA, and Garcia JG. Evans blue dye as a marker of albumin clearance in cultured endothelial monolayer and isolated lung. J Appl Physiol 72: 865-873, 1992.[Abstract/Free Full Text]
- Paulus MJ, Gleason SS, Kennel SJ, Hunsicker PR, and Johnson DK. High resolution X-ray computed tomography: an emerging tool for small animal cancer research. Neoplasia 2: 62-70, 2000.[CrossRef][ISI][Medline]
- Pearce ML, Yamashita J, and Beazell J. Measurement of pulmonary edema. Circ Res 16: 482-488, 1965.[ISI]
- Perry ML, Kayes SG, Barnard JW, and Taylor AE. Effects of phorbol myristate acetate stimulated human leukocytes on rat lung. J Appl Physiol 68: 235-240, 1990.[Abstract/Free Full Text]
- Phelps ME. PET: the merging of biology and imaging into molecular imaging. J Nucl Med 41: 661-681, 2000.[Abstract]
- Pilati CF and Maron MB. A technique to measure the reflection coefficient using endogenous vascular indicators. Microvasc Res 32: 255-260, 1986.[ISI][Medline]
- Pittet JF, Griffiths MJ, Geiser T, Kaminski N, Dalton SL, Huang X, Brown LA, Gotwals PJ, Koteliansky VE, Matthay MA, and Sheppard D. TGF-
is a critical mediator of acute lung injury. J Clin Invest 107: 1537-1544, 2001.[Abstract/Free Full Text]
- Pollick C, Hale SL, and Kloner RA. Echocardiographic and cardiac Doppler assessment of mice. J Am Soc Echocardiogr 8: 602-610, 1995.[Medline]
- Predescu SA, Predescu DN, and Palade GE. Plasmalemmal vesicles function as transcytotic carriers for small proteins in the continuous endothelium. Am J Physiol Heart Circ Physiol 272: H937-H949, 1997.[Abstract/Free Full Text]
- Qiao RL and Bhattacharya J. Segmental barrier properties of the pulmonary microvascular bed. J Appl Physiol 71: 2152-2159, 1991.[Abstract/Free Full Text]
- Razani B, Woodman SE, and Lisanti MP. Caveolae: from cell biology to animal physiology. Pharmacol Rev 54: 431-467, 2002.[Abstract/Free Full Text]
- Reed RK, Townsley MI, Korthuis RJ, and Taylor AE. Analysis of lymphatic protein flux data. V. Unique PS products and sigma dS at low lymph flows. Am J Physiol Heart Circ Physiol 261: H728-H740, 1991.[Abstract/Free Full Text]
- Renkin EM. Cellular and intercellular transport pathways in exchange vessels. Am Rev Respir Dis 146: S28-S31, 1992.[ISI][Medline]
- Reynolds SD, Reynolds PR, Pryhuber GS, Finder JD, and Stripp BR. Secretoglobins SCGB3A1 and SCGB3A2 define secretory cell subsets in mouse and human airways. Am J Respir Crit Care Med 166: 1498-1509, 2001.[CrossRef][ISI]
- Rippe B and Haraldsson B. Transport of macromolecules across microvascular walls: the two-pore theory. Physiol Rev 74: 163-219, 1994.[Abstract/Free Full Text]
- Rippe B, Parker JC, Townsley MI, Mortillaro NA, and Taylor AE. Segmental vascular resistances and compliances in dog lung. J Appl Physiol 62: 1206-1215, 1987.[Abstract/Free Full Text]
- Rippe B, Rosengren BI, Carlsson O, and Venturoli D. Transendothelial transport: the vesicle controversy. J Vasc Res 39: 375-390, 2002.[CrossRef][ISI][Medline]
- Rippe B and Taylor A. NEM and filipin increase albumin transport in lung microvessels. Am J Physiol Heart Circ Physiol 280: H34-H41, 2001.[Abstract/Free Full Text]
- Rippe B, Townsley M, Parker JC, and Taylor AE. Osmotic reflection coefficient for total plasma protein in lung microvessels. J Appl Physiol 58: 436-442, 1985.[Abstract/Free Full Text]
- Rothman JE and Wieland FT. Protein sorting by transport vesicles. Science 272: 227-233, 2003.
- Rutili G, Kvietys P, Martin D, Parker JC, and Taylor AE. Increased pulmonary microvasuclar permeability induced by
-naphthylthiourea. J Appl Physiol 52: 1316-1323, 1982.[Abstract/Free Full Text]
- Saint-Georges L, Van Gorp U, and Maisin JR. Response of mouse lung air-blood barrier to X-irradiation: ultrastructural and stereological analysis. Scanning Microsc 2: 537-543, 1988.[ISI][Medline]
- Schnitzer JE, Allard J, and Oh P. NEM inhibits transcytosis, endocytosis, and capillary permeability: implication of caveolae fusion in endothelia. Am J Physiol Heart Circ Physiol 268: H48-H55, 1995.[Abstract/Free Full Text]
- Schnitzer JE, Liu J, and Oh P. Endothelial caveolae have the molecular transport machinery for vesicle budding, docking, and fusion including VAMP, NSF, SNAP, annexins, and GTPases. J Biol Chem 270: 14399-14404, 1995.[Abstract/Free Full Text]
- Schnitzer JE and Oh P. Albondin-mediated capillary permeability to albumin. Differential role of receptors in endothelial transcytosis and endocytosis of native and modified albumins. J Biol Chem 269: 6072-6082, 1994.[Abstract/Free Full Text]
- Schnitzer JE, Oh P, Pinney E, and Allard J. Filipin-sensitive caveolae-mediated transport in endothelium: reduced transcytosis, scavenger endocytosis, and capillary permeability of select macromolecules. J Cell Biol 127: 1217-1232, 1994.[Abstract]
- Schraufnagel DE and Schmid A. Microvascular casting of the lung: effects of various fixation protocols. J Electron Microsc Tech 8: 185-191, 1988.[ISI][Medline]
- Schubert W, Frank PG, Woodman SE, Hyogo H, Cohen DE, Chow CW, and Lisanti MP. Microvascular hyperpermeability in caveolin-1 (-/-) knock-out mice. Treatment with a specific nitric-oxide synthase inhibitor, L-name, restores normal microvascular permeability in Cav-1 null mice. J Biol Chem 277: 40091-40098, 2002.[Abstract/Free Full Text]
- Scillia P, Delcroix M, Lejeune P, Melot C, Struyven J, Naeije R, and Gevenois PA. Hydrostatic pulmonary edema: evaluation with thin-section CT in dogs. Radiology 211: 161-168, 1999.[Abstract/Free Full Text]
- Scillia P, Kafi SA, Melot C, Keyzer C, Naeije R, and Gevenois PA. Oleic acid-induced lung injury: thin-section CT evaluation in dogs. Radiology 219: 724-731, 2001.[Abstract/Free Full Text]
- Sciuto AM and Moran TS. BHA diet enhances the survival of mice exposed to phosgene: the effect of BHA on glutathione levels in the lung. Inhal Toxicol 11: 855-871, 1999.[CrossRef][ISI][Medline]
- Seibert AF, Thompson WJ, Taylor AE, Wilborn WH, Barnard JW, and Haynes J Jr. Reversal of increased microvascular permeability associated with ischemia-reperfusion: role of cAMP. J Appl Physiol 72: 389-395, 1992.[Abstract/Free Full Text]
- Shibamoto T, Parker JC, Taylor AE, and Townsley MI. Derecruitment of filtration surface area in paraquat-injured isolated dog lungs. J Appl Physiol 68: 1581-1589, 1990.[Abstract/Free Full Text]
- Sibilla S, Tredici S, Porro A, Irace M, Guglielmi M, Nicolini G, Tredici G, Valenza F, and Gattinoni L. Equal increases in respiratory system elastance reflect similar lung damage in experimental ventilator-induced lung injury. Intensive Care Med 28: 196-203, 2002.[CrossRef][ISI][Medline]
- Simionescu N. Transcapillary movement of large molecules in the microvasculature. In: Pulmonary Edema, edited by Fishman AP and Renkin EM. Washington, DC: Am Physiol Soc, 1979, p. 39-52.
- Song Y, Fukuda N, Bai C, Ma T, Matthay MA, and Verkman AS. Role of aquaporins in alveolar fluid clearance in neonatal and adult lung, and in oedema formation following acute lung injury: studies in transgenic aquaporin null mice. J Physiol 525: 771-779, 2000.[Abstract/Free Full Text]
- Steudel W, Ichinose F, Huang PL, Hurford WE, Jones RC, Bevan JA, Fishman MC, and Zapol WM. Pulmonary vasoconstriction and hypertension in mice with targeted disruption of the endothelial nitric oxide synthase (NOS 3) gene. Circ Res 81: 34-41, 1997.[Abstract/Free Full Text]
- Tankersley CG, Rabold R, and Mitzner W. Differential lung mechanics are genetically determined in inbred murine strains. J Appl Physiol 86: 1764-1769, 1999.[Abstract/Free Full Text]
- Taylor AE, Bishop VS, and Guyton AC. Permeability of the alveolar membrane to solutes. Circ Res 16: 353-362, 1965.[ISI]
- Taylor AE and Gaar KAJ. Estimation of equivalent pore radii of pulmonary capillary and alveolar membranes. Am J Physiol 218: 1133-1140, 1970.[Free Full Text]
- Taylor AE and Granger DN. Exchange of macromolecules across the microcirculation. In: Handbook of Physiology, edited by Renkin EM and Michel CC. Bethesda, MD: Am Physiol Soc, 1984, p. 467-520.
- Taylor AE and Parker JC. Pulmonary interstitial spaces and lymphatics. In: Handbook of Physiology. The Respiratory System, edited by Fishman AP and Fisher AB. Bethesda, MD: Am Physiol Society, 1985, sect 3, p. 167-230.
- Terakawa T, Nakata K, Uchida E, Hmano K, Kamio I, Matsushita G, and Tanaka N. Proteome analysis of proteins in bronchoalveolar lavage fluid. Am J Respir Crit Care Med 167: A742, 2003.
- Theodore J, Robin ED, Gaudio R, and Acevedo J. Transalveolar transport of large polar solutes (sucrose, inulin, and dextran). Am J Physiol 229: 989-996, 1975.[Abstract/Free Full Text]
- Tiruppathi C, Finnegan A, and Malik AB. Isolation and characterization of a cell surface albumin-binding protein from vascular endothelial cells. Proc Natl Acad Sci USA 93: 250-254, 1996.[Abstract/Free Full Text]
- Tiruppathi C, Freichel M, Vogel SM, Paria BC, Mehta D, Flockerzi V, and Malik AB. Impairment of store-operated Ca2+ entry in TRPC4(-/-) mice interferes with increase in lung microvascular permeability. Circ Res 91: 70-76, 2002.[Abstract/Free Full Text]
- Townsley MI. Divergent effects of thapsigargin on lung endothelial water and solute permeability (Abstract). Am J Respir Crit Care Med 163: A614, 2001.
- Townsley MI, Bartle SM, Senteno EC, and Barnard JW. Evaluation of leak sites associated with high vascular pressure in the dog lung using vascular corrosion casting (Abstract). FASEB J 3: A1403, 1989.
- Townsley MI, Fu Z, Mathieu-Costello O, and West JB. Pulmonary microvascular permeability. Responses to high vascular pressure after induction of pacing-induced heart failure in dogs. Circ Res 77: 317-325, 1995.[Abstract/Free Full Text]
- Townsley MI, Korthuis RJ, Rippe B, Parker JC, and Taylor AE. Validation of double vascular occlusion method for Pc,i in lung and skeletal muscle. J Appl Physiol 61: 127-132, 1986.[Abstract/Free Full Text]
- Townsley MI, Lim EH, Sahawneh TM, and Song W. Interaction of chemical and high vascular pressure injury in isolated canine lung. J Appl Physiol 69: 1657-1664, 1990.[Abstract/Free Full Text]
- Townsley MI, Parker JC, Korthuis RJ, and Taylor AE. Alterations in hemodynamics and Kf,c during lung mass resection. J Appl Physiol 63: 2460-2466, 1987.[Abstract/Free Full Text]
- Uhlig S, Wollin L, and Wendel A. Contributions of thromboxane and leukotrienes to PAF-induced impairment of lung function in the rat. J Appl Physiol 77: 269-269, 1994.
- Varner AE, Sorkness RL, Kumar A, Kaplan MR, and Lemanske RF Jr. Serial segmental bronchoalveolar lavage in individual rats. J Appl Physiol 87: 1230-1233, 1999.[Abstract/Free Full Text]
- Velazquez M, Haller J, Amundsen T, and Schuster DP. Regional lung water measurements with PET: accuracy, reproducibility, and linearity. J Nucl Med 32: 719-725, 1991.[Abstract]
- Verbrugge SJ, Uhlig S, Neggers SJ, Martin C, Held HD, Haitsma JJ, and Lachmann B. Different ventilation strategies affect lung function but do not increase tumor necrosis factor-
and prostacyclin production in lavaged rat lungs in vivo. Anesthesiology 91: 1834-1843, 1999.[ISI][Medline]
- Verkman AS, Matthay MA, and Song Y. Aquaporin water channels and lung physiology. Am J Physiol Lung Cell Mol Physiol 278: L867-L879, 2000.[Abstract/Free Full Text]
- Vlahakis NE, Schroeder MA, Pagano RE, and Hubmayr RD. Role of deformation-induced lipid trafficking in the prevention of plasma membrane stress failure. Am J Respir Crit Care Med 166: 1282-1289, 2002.[Abstract/Free Full Text]
- Vogel SM, Gao X, Mehta D, Ye RD, John TA, Andrade-Gordon P, Tiruppathi C, and Malik AB. Abrogation of thrombin-induced increase in pulmonary microvascular permeability in PAR-1 knockout mice. Physiol Genomics 4: 137-145, 2000.[Abstract/Free Full Text]
- Vogel SM, Minshall RD, Pilipovic M, Tiruppathi C, and Malik AB. Albumin uptake and transcytosis in endothelial cells in vivo induced by albumin-binding protein. Am J Physiol Lung Cell Mol Physiol 281: L1512-L1522, 2001.[Abstract/Free Full Text]
- Von Bethmann AN, Brasch F, Nusing R, Vogt K, Volk HD, Muller KM, Wendel A, and Uhlig S. Hyperventilation induces release of cytokines from perfused mouse lung. Am J Respir Crit Care Med 157: 263-272, 1998.[ISI][Medline]
- Walters DM, Wills-Karp M, and Mitzner W. Assessment of cellular profile and lung function with repeated bronchoalveolar lavage in individual mice. Physiol Genomics 2: 29-36, 2000.[Abstract/Free Full Text]
- Wang le F, Patel M, Razavi HM, Weicker S, Joseph MG, McCormack DG, and Mehta S. Role of inducible nitric oxide synthase in pulmonary microvascular protein leak in murine sepsis. Am J Respir Crit Care Med 165: 1634-1639, 2002.[Abstract/Free Full Text]
- Ward NS, Waxman AB, Homer RJ, Mantell LL, Einarsson O, Du Y, and Elias JA. Interleukin-6-induced protection in hyperoxic acute lung injury. Am J Respir Cell Mol Biol 22: 535-542, 2000.[Abstract/Free Full Text]
- Waypa GB, Morton CA, Vincent PA, Mahoney JR Jr., Johnston WK III, and Minnear FL. Oxidant-increased endothelial permeability: prevention with phosphodiesterase inhibition vs. cAMP production. J Appl Physiol 88: 835-842, 2000.[Abstract/Free Full Text]
- Weibel ER. Morphometry: stereological theory and practical methods. In: Models of Lung Disease. Microscopy and Structural Methods, edited by Gil J. New York: Dekker, 1990, p. 199-252.
- Weiser MR, Pechet TT, Williams JP, Ma M, Frenette PS, Moore FD, Kobzik L, Hines RO, Wagner DD, Carroll MC, and Hechtman HB. Experimental murine acid aspiration injury is mediated by neutrophils and the alternative complement pathway. J Appl Physiol 83: 1090-1095, 1997.[Abstract/Free Full Text]
- Wiener-Kronish JP and Broaddus VC. Interrelationship of pleural and pulmonary interstitial liquid. Annu Rev Physiol 55: 209-226, 1993.[CrossRef][ISI][Medline]
- Wolf MB, Watson PD, and Scott DR. Integral-mass balance method for determination of solvent drag reflection coefficient. Am J Physiol Heart Circ Physiol 253: H194-H204, 1987.[Abstract/Free Full Text]
- Wu DX, Weibel ER, Bachofen H, and Schurch S. Lung lesions in experimental hydrostatic pulmonary edema: an electron microscopic and morphometric study. Exp Lung Res 21: 711-730, 1995.[ISI][Medline]
- Xu N, Rahman A, Minshall RD, Tiruppathi C, and Malik AB.
2-Integrin blockade driven by E-selectin promoter prevents neutrophil sequestration and lung injury in mice. Circ Res 87: 254-260, 2000.[Abstract/Free Full Text]
- Yoshikawa S, Kayes SG, and Parker JC. Eosinophils increase lung microvascular permeability via the peroxidase-hydrogen peroxide-halide system. Bronchoconstriction and vasoconstriction unaffected by eosinophil peroxidase inhibition. Am Rev Respir Dis 147: 914-920, 1993.[ISI][Medline]
- Yoshikawa S, Parker JC, Sybert C, Eyal FG, and Lausch RN. Ventilator induced lung albumin leak and MIP-2 production independent of TNF's effect in intact mice. Am J Respir Crit Care Med 165: A767-A787, 2002.
- Yoshikawa S, Reynolds SD, and Parker JC. Ventilator induced lung injury detected by plasma levels of Clara cell specific protein in mice. Am J Respir Crit Care Med 167: A775, 2003.