Comprehensive Physiology Wiley Online Library

Modeling of Gas Exchange in the Lungs

Full Article on Wiley Online Library



Abstract

This overview presents the recent progress in our understanding of gas transfer by the lungs during the respiratory cycle and during breath holding. Different phenomena intervene in gas transfer, convection and diffusion in the gas, dissolution, diffusion across the alveolar‐capillary membrane, diffusion across blood plasma, and finally diffusion and reaction with hemoglobin inside blood cells. The different gases, O2, CO, and NO, have very different reaction times with hemoglobin ranging from a few microseconds to tens of milliseconds. This is leading to different outcomes.

For O2, the solutions to the coupled nonlinear gas and blood equations are obtained at the acinus level. They include the fact that the acinar internal ventilation is strongly heterogeneous due to the arborescent structure. Also, in the dynamic calculation, one takes care of the delay between the start of inhalation and arrival of fresh air in the acinus. This “dead” time is the dynamic equivalent of the dead space ventilation.

The question of the dependence of Vo2 on ventilation and perfusion takes a different form. The results show that Vo2 is not only a function of the ventilation/perfusion ratio but also depends on the variables: acinar ventilation VEac and perfusion Qac. The ratio VEac/Qac roughly determines arterial O2 saturation and arterial and alveolar O2 partial pressure.

The classic Roughton‐Forster interpretation of DLCO (separation between independent membrane and blood resistance) was a mathematical conjecture. It was shown recently that this conjecture was violated. This article presents an alternative interpretation that uses time concepts instead of resistance. © 2021 American Physiological Society. Compr Physiol 11:1289‐1314, 2021.

Figure 1. Figure 1. Penetrations of CO and NO obtained by 2D numerical solutions of Eq. 7 in the RBC morphologies obtained by Weibel 84 (A, alveolar air; EC, RBC; P, plasma; T, tissue barrier). The colors correspond to the concentrations of CO or NO. The dark red regions illustrate the active volumes where the majority of the transfer occurs. The penetration from the surface is approximately represented by the thickness of the dark red region. In the cluster shown here, the region active for CO transfer is the majority of the cluster volume, whereas the region active for NO transfer is the thin dark red layer.
Figure 2. Figure 2. Qualitative schematic of the problem that is solved numerically here: profile of O2 partial pressures along the airway tract during inspiration and along the perfusion pathway from the pulmonary artery to the capillary units and the pulmonary vein. The gas‐exchange units corresponding to the capillary units are perfused in parallel but ventilated in series. Courtesy of E. R. Weibel.
Figure 3. Figure 3. Due to the short transverse diffusion time, the numerical problem can be mapped on tree made of one‐dimensional branches. The x coordinate follows the tree branches.
Figure 4. Figure 4. Time evolution of blood saturation for a diffusion time across membrane and plasma of 1 ms and for two values of the oxygen partial pressure in the gas: Pg = 70 mmHg and 100 mmHg. Also shown in blue is the saturation evolution predicted by Whiteley 90 from a finite element calculation. The final saturation value is close to 0.96 for Pg(x,t) = 100 mmHg. The arrows are used to define the approximate saturation time ts. See Kang et al. (2015) for details.
Figure 5. Figure 5. Results of the dynamic oxygen uptake model for rest and heavy exercise after Kang et al. 42. (A) Spatiotemporal distribution of the local alveolar O2 partial pressure in the acinus. (B) Spatiotemporal distribution of the local transfer. For visibility, the spatial axis is inversed as indicated by the arrow. Data are computed for a situation where the time of arrival (dead time) is 0.7 s at rest and 0.15 s at exercise.
Figure 6. Figure 6. Schematic representation of the DOT model used by Kang et al. Given the average morphology of an acinus, the dynamic model relates the four principal respiratory determinants (VE, PIO2, Q, and PvO2) to the four outcome values of O2 transfer, namely, Vo2, SaO2, Pao2, and PAO2. The subscript ac indicates that the values are for an acinus.
Figure 7. Figure 7. Functioning of single standard acinus at rest for a range of ventilation and perfusion computed by the DOT model. Left and bottom axes of each map represent acinar ventilation and perfusion values, respectively. Right and top axes indicate global values of ventilation and perfusion, respectively, calculated by multiplying the acinar ventilation and perfusion by the number of acini corresponding to a “standard” subject with TLC = 6.2 liter and dead space volume VD = 0.17 liter. Some labels are omitted due to limited space. The blue arrows in the top‐left map illustrate the inverse approach for retrieving Qac from VEac and Vo2ac (see text). The red lines in these figures represent (isoventilation/perfusion) ratio lines, with VEac/Qac = 1 (solid); VEac/Qac = 1.5 (dotted); and VEac/Qac = 0.5 (broken).
Figure 8. Figure 8. Functioning of single standard acinus at peak exercise for a range of ventilation and perfusion computed by the DOT model. Left and bottom axes of each map represent acinar ventilation and perfusion values, respectively. Right and top axes indicate global values of ventilation and perfusion, respectively, calculated by multiplying the acinar ventilation and perfusion by the number of acini corresponding to a standard subject with TLC = 6.2 liter and dead space volume VD = 0.17 liter. Some labels are omitted due to limited space. The red lines represent isoventilation over perfusion ratio lines, with VEac/Qac = 1 (solid), VEac/Qac = 1.5 (dotted), and VEac/Qac = 2 (broken).
Figure 9. Figure 9. Comparison between individual cardiac output values deduced from the inverse approach and their corresponding values deduced from Fick principle. Each point corresponds to a given individual. Red circles are for rest, and green triangles for exercise. Here, the predicted values have been calculated using, in addition to the four primary determinants, two secondary individual determinants, namely, Hb concentration and individual P50.
Figure 10. Figure 10. DOT model predictions of high‐altitude maximal oxygen uptake (Vo2max), red curve, and its comparison to experimental data.
Figure 11. Figure 11. Gas exchange under significant diffusion disturbance due to increase in membrane thickness as predicted by the model. Top: for rest. Bottom: for peak exercise.
Figure 12. Figure 12. Dependence of oxygen transfer Vo2, saturation SaO2, mean alveolar pressure PAO2, and arterial pressure Pao2 on the arrival (dead) time at which the oxygenated fresh air reaches an acinus at constant VE = 7.5 liter/min (computed for PvO2 = 40 mmHg and Q = 5 liter/min). Each value was normalized to its value at zero arrival time tarr = 0 s. This is why the PA and Pa curves coincide. The vertical dotted line represents a “normal” reference condition of tarr = 0.7 s. Note that the PAO2 and Pao2 curves are practically the same because the ordinates are in relative units, and PAO2‐Pao2 keeps small.
Figure 13. Figure 13. Two‐dimensional (2D) illustration of the distribution of volumes of fresh air crossing generation 12 in the Florens et al. model 17. There are 212 squares, each of which represents one airway at generation 12. The colors correspond to the logarithm of the distributed volume of fresh air expressed in milliliters. The black squares correspond to airways that are too small (diameter smaller than 0.5 mm) to belong to the tracheobronchial tree. Note the logarithmic color scale.
Figure 14. Figure 14. Time evolution of the spatial distribution of CO concentration. t represents the time after the RBC enters in diffusive contact with the gas. The membrane thickness is chosen to be 1 μm. Concentrations are normalized to the concentration at the gas‐membrane interface Cg. The horizontal line represents the membrane‐plasma interface. The bottom figure gives the rate of transfer as a function of time. It reaches a steady state after 2 to 3 ms.
Figure 15. Figure 15. Three simplified geometries of the gas‐membrane‐plasma‐RBC structure. Top A: Horizontal cut of a planar parallel structure of a symmetrical gas‐membrane‐plasma‐RBC geometry. This is an artificial case where the alveolar‐capillary membrane, the plasma layer, and the blood cells are supposed to be planar with thicknesses of LM, LP, and LRBC, respectively. The dotted line at the bottom indicates the plane of symmetry. Top B: Quarter of an axial planar cross section of spherical RBC inside a cylindrical capillary shown as the bottom revolution axis. Top C: Quarter of an axial planar cross section of biconcave RBC in a cylindrical capillary. Bottom: 3D illustrations of each situation.
Figure 16. Figure 16. Spatial distribution of CO and NO concentrations in the planar system. Concentrations at the gas membrane interface have been normalized to 1. For simplicity, the same diffusivity is applied to the entire system. (Computed for DCO = 3.26 × 10−5 cm2/s; DNO = 3.18 × 10−5 cm2/s; τCO = 0.5 ms; τNO = 0.005 ms; LM = 1 μm; LP = 2.31 μm; LRBC = 3.08 μm. See below for these values.)
Figure 17. Figure 17. Results of the exact calculation of 1/DLCO as a function of the reaction time τ (τ is proportional to PO2 for PO2 > 80 mmHg) for the planar system of Figure 16.
Figure 18. Figure 18. Comparison of the planar model prediction using Eq. 34 and experimental data from Forster et al. 21. Values of 1/DLCO are rescaled using the values 1/DLCO (Ext.Pao2 → 0 mmHg) (Eq. 41).


Figure 1. Penetrations of CO and NO obtained by 2D numerical solutions of Eq. 7 in the RBC morphologies obtained by Weibel 84 (A, alveolar air; EC, RBC; P, plasma; T, tissue barrier). The colors correspond to the concentrations of CO or NO. The dark red regions illustrate the active volumes where the majority of the transfer occurs. The penetration from the surface is approximately represented by the thickness of the dark red region. In the cluster shown here, the region active for CO transfer is the majority of the cluster volume, whereas the region active for NO transfer is the thin dark red layer.


Figure 2. Qualitative schematic of the problem that is solved numerically here: profile of O2 partial pressures along the airway tract during inspiration and along the perfusion pathway from the pulmonary artery to the capillary units and the pulmonary vein. The gas‐exchange units corresponding to the capillary units are perfused in parallel but ventilated in series. Courtesy of E. R. Weibel.


Figure 3. Due to the short transverse diffusion time, the numerical problem can be mapped on tree made of one‐dimensional branches. The x coordinate follows the tree branches.


Figure 4. Time evolution of blood saturation for a diffusion time across membrane and plasma of 1 ms and for two values of the oxygen partial pressure in the gas: Pg = 70 mmHg and 100 mmHg. Also shown in blue is the saturation evolution predicted by Whiteley 90 from a finite element calculation. The final saturation value is close to 0.96 for Pg(x,t) = 100 mmHg. The arrows are used to define the approximate saturation time ts. See Kang et al. (2015) for details.


Figure 5. Results of the dynamic oxygen uptake model for rest and heavy exercise after Kang et al. 42. (A) Spatiotemporal distribution of the local alveolar O2 partial pressure in the acinus. (B) Spatiotemporal distribution of the local transfer. For visibility, the spatial axis is inversed as indicated by the arrow. Data are computed for a situation where the time of arrival (dead time) is 0.7 s at rest and 0.15 s at exercise.


Figure 6. Schematic representation of the DOT model used by Kang et al. Given the average morphology of an acinus, the dynamic model relates the four principal respiratory determinants (VE, PIO2, Q, and PvO2) to the four outcome values of O2 transfer, namely, Vo2, SaO2, Pao2, and PAO2. The subscript ac indicates that the values are for an acinus.


Figure 7. Functioning of single standard acinus at rest for a range of ventilation and perfusion computed by the DOT model. Left and bottom axes of each map represent acinar ventilation and perfusion values, respectively. Right and top axes indicate global values of ventilation and perfusion, respectively, calculated by multiplying the acinar ventilation and perfusion by the number of acini corresponding to a “standard” subject with TLC = 6.2 liter and dead space volume VD = 0.17 liter. Some labels are omitted due to limited space. The blue arrows in the top‐left map illustrate the inverse approach for retrieving Qac from VEac and Vo2ac (see text). The red lines in these figures represent (isoventilation/perfusion) ratio lines, with VEac/Qac = 1 (solid); VEac/Qac = 1.5 (dotted); and VEac/Qac = 0.5 (broken).


Figure 8. Functioning of single standard acinus at peak exercise for a range of ventilation and perfusion computed by the DOT model. Left and bottom axes of each map represent acinar ventilation and perfusion values, respectively. Right and top axes indicate global values of ventilation and perfusion, respectively, calculated by multiplying the acinar ventilation and perfusion by the number of acini corresponding to a standard subject with TLC = 6.2 liter and dead space volume VD = 0.17 liter. Some labels are omitted due to limited space. The red lines represent isoventilation over perfusion ratio lines, with VEac/Qac = 1 (solid), VEac/Qac = 1.5 (dotted), and VEac/Qac = 2 (broken).


Figure 9. Comparison between individual cardiac output values deduced from the inverse approach and their corresponding values deduced from Fick principle. Each point corresponds to a given individual. Red circles are for rest, and green triangles for exercise. Here, the predicted values have been calculated using, in addition to the four primary determinants, two secondary individual determinants, namely, Hb concentration and individual P50.


Figure 10. DOT model predictions of high‐altitude maximal oxygen uptake (Vo2max), red curve, and its comparison to experimental data.


Figure 11. Gas exchange under significant diffusion disturbance due to increase in membrane thickness as predicted by the model. Top: for rest. Bottom: for peak exercise.


Figure 12. Dependence of oxygen transfer Vo2, saturation SaO2, mean alveolar pressure PAO2, and arterial pressure Pao2 on the arrival (dead) time at which the oxygenated fresh air reaches an acinus at constant VE = 7.5 liter/min (computed for PvO2 = 40 mmHg and Q = 5 liter/min). Each value was normalized to its value at zero arrival time tarr = 0 s. This is why the PA and Pa curves coincide. The vertical dotted line represents a “normal” reference condition of tarr = 0.7 s. Note that the PAO2 and Pao2 curves are practically the same because the ordinates are in relative units, and PAO2‐Pao2 keeps small.


Figure 13. Two‐dimensional (2D) illustration of the distribution of volumes of fresh air crossing generation 12 in the Florens et al. model 17. There are 212 squares, each of which represents one airway at generation 12. The colors correspond to the logarithm of the distributed volume of fresh air expressed in milliliters. The black squares correspond to airways that are too small (diameter smaller than 0.5 mm) to belong to the tracheobronchial tree. Note the logarithmic color scale.


Figure 14. Time evolution of the spatial distribution of CO concentration. t represents the time after the RBC enters in diffusive contact with the gas. The membrane thickness is chosen to be 1 μm. Concentrations are normalized to the concentration at the gas‐membrane interface Cg. The horizontal line represents the membrane‐plasma interface. The bottom figure gives the rate of transfer as a function of time. It reaches a steady state after 2 to 3 ms.


Figure 15. Three simplified geometries of the gas‐membrane‐plasma‐RBC structure. Top A: Horizontal cut of a planar parallel structure of a symmetrical gas‐membrane‐plasma‐RBC geometry. This is an artificial case where the alveolar‐capillary membrane, the plasma layer, and the blood cells are supposed to be planar with thicknesses of LM, LP, and LRBC, respectively. The dotted line at the bottom indicates the plane of symmetry. Top B: Quarter of an axial planar cross section of spherical RBC inside a cylindrical capillary shown as the bottom revolution axis. Top C: Quarter of an axial planar cross section of biconcave RBC in a cylindrical capillary. Bottom: 3D illustrations of each situation.


Figure 16. Spatial distribution of CO and NO concentrations in the planar system. Concentrations at the gas membrane interface have been normalized to 1. For simplicity, the same diffusivity is applied to the entire system. (Computed for DCO = 3.26 × 10−5 cm2/s; DNO = 3.18 × 10−5 cm2/s; τCO = 0.5 ms; τNO = 0.005 ms; LM = 1 μm; LP = 2.31 μm; LRBC = 3.08 μm. See below for these values.)


Figure 17. Results of the exact calculation of 1/DLCO as a function of the reaction time τ (τ is proportional to PO2 for PO2 > 80 mmHg) for the planar system of Figure 16.


Figure 18. Comparison of the planar model prediction using Eq. 34 and experimental data from Forster et al. 21. Values of 1/DLCO are rescaled using the values 1/DLCO (Ext.Pao2 → 0 mmHg) (Eq. 41).
References
 1.Aguilaniu B, Maitre J, Glenet S, Gegout‐Petit A, Guenard H. European reference equations for CO and NO lung transfer. Eur Respir J 31: 1091‐1097, 2008.
 2.Angelov B, Mladenov IM. On the geometry of the red blood cell. In: Mladenov IM, Naber GL, editors. Geometry, Integrability and Quantization. Sofia: Coral Press, 2000, p. 27‐46.
 3.Anthonisen NR, Fleetham JA. Ventilation: Total, alveolar, and dead space. In: R Terjung (Ed.). Comprehensive Physiology, 2012. DOI:10.1002/cphy.cp030407.
 4.Beck KC, Johnson BD, Olson TP, Wilson TA. Ventilation‐perfusion distribution in normal subjects. J Appl Physiol 113 (6): 872‐877, 2012.
 5.Borland CDR, Higenbottam TW. A simultaneous single breath measurement of pulmonary diffusing capacity with nitric oxide and carbon monoxide. Eur Respir J 2: 56‐63, 1989.
 6.Borland CDR, Cox Y. Effect of varying alveolar oxygen partial pressure on diffusing capacity for nitric oxide and carbon monoxide, membrane diffusing capacity and lung capillary blood volume. Clin Sci 81: 759‐765, 1991.
 7.Borland CDR, Dunningham H, Bottrill F, Vuylsteke A, Yilmaz C, Hsia CCW. Significant blood resistance to nitric oxide transfer in the lung. J Appl Physiol 108: 1052‐1060, 2010.
 8.Borland CDR, Bottrill F, Jones A, Sparkes C, Vuylsteke A. The significant blood resistance to lung nitric oxide transfer lies within the red cell. J Appl Physiol 116: 32‐41, 2014.
 9.Borland C, Moggridge G, Patel R, Patel S, Zhu Q, Vuylsteke A. Permeability and diffusivity of nitric oxide in human plasma and red cells. Nitric Oxide 78: 51‐59, 2018.
 10.Chakraborty S, Balakotaiah V, Bidani A. Diffusing capacity reexamined: Relative roles of diffusion and chemical reaction in red cell uptake of O2, CO, CO2, and NO. J Appl Physiol 97: 2284‐2302, 2004.
 11.Checchia PA, Bronicki RA, Goldstein B. Review of inhaled nitric oxide in the pediatric cardiac surgery setting. Pediatr Cardiol 33: 493‐505, 2012.
 12.Coffman KE, Taylor BJ, Carlson AR, Wentz RJ, Johnson BD. Optimizing the calculation of DM CO and VC via the single breath single oxygen tension DLCO/NO method. Respir Physiol Neurobiol 221: 19‐29, 2016.
 13.Comroe JH Jr. Pulmonary diffusing capacity for carbon monoxide (DLCO). Am Rev Respir Dis 111: 225‐228, 1975.
 14.Cotes JE, Chinn DJ, Lung Function MMR. Physiology, Measurement and Application in Medicine (6th ed). Malden, MA: Blackwell Publishing Ltd, 2006, p. 224‐229.
 15.Dash RK, Korman B, Bassingthwaighte J. Simple accurate mathematical models of blood HbO2 and HbCO2 dissociation curves at varied physiological conditions: Evaluation and comparison with other models. Eur J Appl Physiol 16 (1): 97‐113, 2016. DOI: 10.1007/s00421‐015‐3228‐3.
 16.Department of Army US. Altitude acclimatization and illness management. Tech Bull Med: 505, 2010.
 17.Florens M, Sapoval B, Filoche M. An anatomical and functional model of the human tracheobronchial tree. J Appl Physiol 110: 756‐763, 2011.
 18.Florens M, Sapoval B, Filoche M. Optimal branching asymmetry of hydrodynamic pulsatile trees. Phys Rev Lett 106: 178104, 2011.
 19.Florens M, Sapoval B, Filoche M. The optimal branching asymmetry of a bidirectional distribution tree. Comput Phys Commun 182 (9): 1932‐1936, 2011.
 20.Formenti F, Bommakanti N, Chen R, Cronin JN, McPeak H, Holopherne‐Doran D, Hedenstierna G, Hahn CE, Larsson A, Farmery AD. Respiratory oscillations in alveolar oxygen tension measured in arterial blood. Sci Rep 7 (1): 7499, 2017.
 21.Forster RE, Roughton FJW, Cander L, Briscoe WA, Kreuzer F. Apparent pulmonary diffusing capacity for CO at varying alveolar O2 tensions. J Appl Physiol 11: 277‐289, 1957.
 22.Foucquier A. Dynamique du transport et du transfert de l'oxygène au sein de l'acinus pulmonaire humain. Thèse de Doctorat de l'Ecole polytechnique, 108‐121, 2010. http://pastel.archives‐ouvertes.fr/index.php?halsid=7ah4v9djubjgh1c9tvert860o5&action_todo=homeunder the name Foucquier.
 23.Foucquier A, Filoche M, Moreira AA, Andrade JS Jr, Arbia G, Sapoval B. A first principles calculation of the oxygen uptake in the human pulmonary acinus at maximal exercise. Respir Physiol Neurobiol 185: 625‐638, 2013.
 24.Gehr P, Bachofen M, Weibel ER. The normal human lung: Ultrastructure and morphometric estimation of diffusion capacity. Respir Physiol 32: 121‐140, 1978.
 25.Gibson QH. The kinetics of reactions between haemoglobin and gases. Prog Biophys Chem 9: 1‐53, 1959.
 26.Glénet SN, De Bisschop C, Vargas F, Guénard HJP. Deciphering the nitric oxide to carbon monoxide lung transfer ratio: Physiological implications. J Physiol 582: 767‐775, 2007.
 27.Glenny RW, Robertson HT. Spatial distribution of ventilation and perfusion: Mechanisms and regulation. Compr Physiol 1: 375‐395, 2011.
 28.Graham BL, Brusasco V, Burgos F, Cooper BG, Jensen R, Kendrick A, MacIntyre NR, Thompson BR, Wanger J. ERS/ATS standards for single‐breath carbon monoxide uptake in the lung. Eur Respir J 49 (1): 1600016, 2017. DOI: 10.1183/13993003.00016‐2016.
 29.Guénard H, Varène N, Vaida P. Determination of the lung capillary blood volume and membrane diffusing capacity in man by the measurements of NO and CO transfer. Respir Physiol 70: 113‐120, 1987.
 30.Guénard HJ, Martinot JB, Martin S, Maury B, Lalande S, Kays C. In vivo estimates of NO and CO conductance for haemoglobin and for lung transfer in humans. Respir Physiol Neurobiol 228: 1‐8, 2016.
 31.Haefeli‐Bleuer B, Weibel ER. Morphometry of the human pulmonary acinus. Anat Rec 220: 401‐414, 1988.
 32.Hahn CE, Farmer AD. Gas exchange modelling: No more gills please. Br J Anaesth 91: 2‐15, 2003.
 33.Hofemeier P, Shachar‐Berman L, Tenenbaum‐Katan J, Filoche M, Sznitman J. Unsteady diffusional screening in 3D pulmonary acinar structures: From infancy to adulthood. J Biomech 49 (11): 2193‐2200, 2016.
 34.Hogg JC, Coxson HO, Brumwell ML, Beyers N, Doerschuk CM, MacNee W, Wiggs BR. Erythrocyte and polymorphonuclear cell transit time and concentration in human pulmonary capillaries. J Appl Physiol 77 (4): 1795‐1800, 1994.
 35.Horsfield K, Dart G, Olson DE, Filley GF, Cumming G. Models of the human bronchial tree. J Appl Physiol 31: 207‐217, 1971.
 36.Hughes JMB, Pride NB eds. Lung function tests: Physiological principles and clinical applications. In: Bailliere Tindall, 1999.
 37.Hughes JMB, Bates DV. Historical review: The carbon monoxide diffusing capacity (DLCO) and its membrane (Dm) and red cell (Θ·Vc) components. Respir Physiol Neurobiol 138: 115‐142, 2003.
 38.Hughes JMB, Pride NB. Examination of the carbon monoxide diffusing capacity (DLCO) in relation to its KCO and VA components. Am J Respir Crit Care Med 186: 132‐139, 2012.
 39.Hughes JMB, van der Lee I. The TLNO/TLCO ratio in pulmonary function test interpretation. Eur Respir J 41: 453‐461, 2013.
 40.Hughes JMB, Borland CD. The centenary (2015) of the transfer factor for carbon monoxide (TLCO): Marie Krogh's legacy. Thorax 70: 391‐394, 2015.
 41.Hughes JMB, Dinh‐Xuan AT. The DLNO/DLCO ratio: Physiological significance and clinical implications. Respir Physiol Neurobiol 241: 17‐22, 2017. DOI: 10.1016/j.resp.2017.01.002.
 42.Kang MY, Katz I, Sapoval B. A new approach to the dynamics of oxygen capture by the human lung. Respir Physiol Neurobiol 205: 109‐119, 2015. DOI: 10.1016/j.resp.2014.11.001.
 43.Kang MY, Sapoval B. Time‐based understanding of DLCO and DLNO. Respir Physiol Neurobiol 225: 48‐59, 2016.
 44.Kang MY, Sapoval B. Prediction of maximal oxygen uptake at high altitude. Eur Respir J 48: PA1583, 2016. DOI: 10.1183/13993003.congress‐2016.PA1583.
 45.Kang MY, Grebenkov D, Guénard H, Katz I, Sapoval B. The Roughton‐Forster equation for DLCO and DLNO re‐examined. Respir Physiol Neurobiol 241: 62‐71, 2017.
 46.Kang MY, Guénard H, Sapoval B. Diffusion reaction of carbon monoxide in the human lung. Phys Rev Lett 119 (7): 078101, 2017.
 47.Kang MY, Hua‐Huy T, Günther S, Aubourg F, Dinh‐Xuan AT, Sapoval B. Individual modeling of oxygen capture by the human lungs. Respir Physiol Neurobiol 270: 103256, 2019.
 48.Kanner RE, Crapo RO. The relationship between alveolar oxygen tension and the single‐breath carbon monoxide diffusing capacity. Am Rev Respir Dis 133: 676‐678, 1986.
 49.Katz I, Pichelin M, Montesantos S, Kang MY, Sapoval B, Zhu K, Thevenin CP, McCoy R, Martin AR, Caillibotte G. An in silico analysis of oxygen uptake of a mild COPD patient during rest and exercise using a portable oxygen concentrator. Int J Chron Obstruct Pulmon Dis 11: 2427‐2434, 2016.
 50.Kitaoka H, Takaki R, Suki B. A three‐dimensional model of the human airway tree. J Appl Physiol 87: 2207‐2217, 1999.
 51.Krogh M. The diffusion of gases through the lung of man. J Physiol 49: 271‐296, 1915.
 52.Levitt DG, Levitt MD. Carbon monoxide: A critical quantitative analysis and review of the extent and limitations of its second messenger function. Clin Pharm 7: 37, 2015.
 53.MacIntyre N, Crapo RO, Viegi G, Johnson DC, van der Grinten CPM, Brusasco V, Burgos F, Casaburi R, Coates A, Enright P, Gustafsson P, Hankinson J, Jensen R, McKay R, Miller MR, Navajas D, Pedersen OF, Pellegrino R, Wanger J. Standardisation of the single‐breath determination of carbon monoxide uptake in the lung. Eur Respir J 26: 720‐735, 2005.
 54.Majumdar A, Alencar AM, Buldyrev SV, Hantos Z, Lutchen KR, Stanley HE, Suki B. Relating airway diameter distributions to regular branching asymmetry in the lung. Phys Rev Lett 95: 168101, 2005.
 55.Martin S, Maury B. Modeling of the oxygen transfer in the respiratory process. Esaim Math Model Numer Anal 47: 935‐960, 2013.
 56.Martinot JB, Guénard HJP. TLNO/TLCO ratio is not the end of the road. Eur Respir J 43: 1535‐1536, 2014.
 57.McWhirter JL, Noguchi H, Gompper G. Flow‐induced clustering and alignment of vesicles and red blood cells in microcapillaries. Proc Natl Acad Sci U S A 106: 6039‐6043, 2009.
 58.Meakin P, Sapoval B. Random‐walk simulation of the response of irregular or fractal interfaces and membranes. Phys Rev A 43 (6): 2993, 1991.
 59.Montesantos S, Katz I, Pichelin M, Caillibotte G. The creation and statistical evaluation of a deterministic model of the human bronchial tree from HRCT images. PLoS One 11 (12): e0168026, 2016. DOI: 10.1371/journal.pone.0168026.
 60.Motterlini R, Otterbein LE. The therapeutic potential of carbon monoxide. Nat Rev Drug Discov 9: 728‐743, 2010.
 61.Noguchi N, Gompper G. Shape transitions of fluid vesicles and red blood cells in capillary flows. Proc Natl Acad Sci U S A 02 (40): 14159‐14164, 2005.
 62.Paiva M, Engel LA. Model analysis of intra‐acinar gas exchange. Respir Physiol 62: 257‐272, 1985.
 63.Palombo M, Gabrielli A, Servedio VDP, Ruocco G, Capuani S. Structural disorder and anomalous diffusion in random packing of spheres. Sci Rep 3: 2631.
 64.Phillips CG, Kaye SR. On the asymmetry of bifurcations in the bronchial tree. Respir Physiol 107: 85‐98, 1997.
 65.Raabe OG, Yeh HC, Schum GM, Phalen RF. Tracheobronchial geometry: Human, dog, rat, hamster. Tech Rep Publ No LF‐53: 23, 1976.
 66.Reeves RB, Park HK. CO uptake kinetics of red cells and CO diffusing capacity. Respir Physiol 88: 1‐21, 1992.
 67.Richalet JP, Herry JP. Médecine De Montagne: Alpinisme Et Sports De Montagne, Paris: Elsevier Masson, 2006.
 68.Riley RL, Cournand A. Ideal' alveolar air and the analysis of ventilation‐perfusion relationships in the lungs. J Appl Physiol 1 (12): 825‐847, 1949.
 69.Roughton FJW, Forster RE. Relative importance of diffusion and chemical reaction rates in determining rate of exchange of gases in the human lung, with special reference to true diffusing capacity of pulmonary membrane and volume of blood in the lung capillaries. J Appl Physiol 11: 290‐302, 1957.
 70.Sakai H, Sato A, Masuda K, Takeoka S, Tsuchida E. Encapsulation of concentrated hemoglobin solution in phospholipid vesicles retards the reaction with NO, but Not CO by intracellular diffusion barrier. J Biol Chem 283: 1508‐1517, 2008.
 71.Sapoval B. General formulation of Laplacian transfer across irregular surfaces. Phys Rev Lett 73: 3314‐3316, 1994.
 72.Sapoval B, Weibel ER, Filoche M. Diffusion screening, acinus size and optimal design of mammalian lungs. In: Losa GA, Merlini D, Nonnenmacher TF, Weibel ER, editors. Fractals in Biology and Medicine. Basel: Birkhäuser Verlag, vol. IV, 2002, p. 25‐38.
 73.Sapoval B, Filoche M, Weibel ER. Smaller is better but not too small: A physical scale for the design of the mammalian pulmonary acinus. Proc Natl Acad Sci U S A 99: 10411‐10416, 2002.
 74.Scheid P, Piiper J. Intrapulmonary gas mixing and stratification. In: West JB, editor. Pulmonary Gas Exchange. New York, San Francisco: Academic Press, vol. I, 1980, p. 87‐130.
 75.Scheid P, Piiper J. Intrapulmonary gas mixing and stratification. Pulm Gas Exch 1: 87‐130, 1980.
 76.Scheid P, Piiper J. Blood gas equilibration in lungs and pulmonary diffusing capacity. In: Chang HK, editor. Lung Biology in Health and Disease, Respiratory Physiology, vol. 40, 1980, p. 453‐497.
 77.Swan AJ, Tawhai MH. Evidence for minimal oxygen heterogeneity in the healthy human pulmonary acinus. J Appl Physiol 110: 528‐537, 2011.
 78.Tabuchi A, Styp‐Rekowska B, Slutsky AS, Wagner PD, Pries AR, Kuebler WM. Precapillary oxygenation contributes relevantly to gas exchange in the intact lung. Am J Respir Crit Care Med 188: 474‐481, 2013.
 79.Turgeon ML. Clinical Hematology: Theory and Procedures. Philadelphia, PA: Lippincott Williams & Wilkins, 2004, p. 100.
 80.van der Lee I, Zanen P, Biesma DH, van den Bosch JM. The effect of red cell transfusion on nitric oxide diffusing capacity. Respiration 72: 512‐516, 2005.
 81.van der Lee I, Zanen P, Grutters JC, Snijder RJ, van den Bosch JM. Diffusing capacity for nitric oxide and carbon monoxide in patients with diffuse parenchymal lung disease and pulmonary arterial hypertension. Chest 129: 378‐383, 2006.
 82.Wagner PD, Saltzman HA, West JB. Measurement of continuous distributions of ventilation‐perfusion ratios: Theory. J Appl Physiol 36 (5): 588‐599, 1974.
 83.Wagner PD. Ventilation‐perfusion relationships. Annu Rev Physiol 42 (1): 235‐247, 1980.
 84.Weibel ER. The Pathway for Oxygen. Cambridge, MA: Harvard University Press, 1984.
 85.Weibel ER. Lung morphometry: The link between structure and function. Cell Tissue Res 367: 413‐426, 2017.
 86.Weibel ER, Sapoval B, Filoche M. Design of peripheral airways for efficient gas exchange. Respir Physiol Neurobiol 148: 3‐21, 2005.
 87.West JB. A century of pulmonary gas exchange. Am J Respir Crit Care Med 169: 897‐902, 2004.
 88.West JB. Respiratory Physiology: The Essentials (8th ed). Baltimore, PA: Wolters Kluwer/Lippincott Williams & Wilkins, 2008.
 89.West JB. History of respiratory gas exchange. Compr Physiol 1: 1509‐1523, 2011.
 90.Whiteley JP. Some factors affecting pulmonary oxygen transport. Math Biosci 202: 115‐132, 2006.
 91.Wise DL, Houghton G. Diffusion coefficient of neon, krypton, xenon, carbon monoxide and nitric oxide in water at 10‐60 °C. Chem Eng Sci 23: 1211‐1216, 1968.
 92.Yamaguchi K, Nguyen‐Phu D, Scheid P, Piiper J. Kinetics of O2 uptake and release by human erythrocytes studied by a stopped‐flow technique. J Appl Physiol 58: 1215‐1224, 1985.
 93.Zacharia IG, Deen WM. Diffusivity and solubility of nitric oxide in water and saline. Ann Biomed Eng 33: 214‐222, 2005.
 94.Zavorsky GS, Cao J, Murias JM. Reference values of pulmonary diffusing capacity for nitric oxide in an adult population. Nitric Oxide 18: 70‐79, 2008.
 95.Zavorsky GS, Hsia CC, Hughes JM, Borland CD, Guénard H, van der Lee I, Steenbruggen I, Naeije R, Cao J, Dinh‐Xuan AT. Standardisation and application of the single‐breath determination of nitric oxide uptake in the lung. Eur Respir J 49: 1600962, 2017. DOI: 10.1183/13993003.00962‐2016.

Contact Editor

Submit a note to the editor about this article by filling in the form below.

* Required Field

How to Cite

Bernard Sapoval, Min‐Yeong Kang, Anh Tuan Dinh‐Xuan. Modeling of Gas Exchange in the Lungs. Compr Physiol 2020, 11: 1289-1314. doi: 10.1002/cphy.c190019