Comprehensive Physiology Wiley Online Library

Hyperbaric Conditions

Full Article on Wiley Online Library



Abstract

Exposure to elevated ambient pressure (hyperbaric conditions) occurs most commonly in underwater diving, during which respired gas density and partial pressures, work of breathing, and physiological dead space are all increased. There is a tendency toward hypercapnia during diving, with several potential causes. Most importantly, there may be reduced responsiveness of the respiratory controller to rising arterial CO2, leading to hypoventilation and CO2 retention. Contributory factors may include elevated arterial PO2, inert gas narcosis and an innate (but variable) tendency of the respiratory controller to sacrifice tight control of arterial CO2 when work of breathing increases. Oxygen is usually breathed at elevated partial pressure under hyperbaric conditions. Oxygen breathing at modest hyperbaric pressure is used therapeutically in hyperbaric chambers to increase arterial carriage of oxygen and diffusion into tissues. However, to avoid cerebral and pulmonary oxygen toxicity during underwater diving, both the magnitude and duration of oxygen exposure must be managed. Therefore, most underwater diving is conducted breathing mixtures of oxygen and inert gases such as nitrogen or helium, often simply air. At hyperbaric pressure, tissues equilibrate over time with high inspired inert gas partial pressure. Subsequent decompression may reduce ambient pressure below the sum of tissue gas partial pressures (supersaturation) which can result in tissue gas bubble formation and potential injury (decompression sickness). Risk of decompression sickness is minimized by scheduling time at depth and decompression rate to limit tissue supersaturation or size and profusion of bubbles in accord with models of tissue gas kinetics and bubble formation and growth. © 2011 American Physiological Society. Compr Physiol 1:163‐201, 2011.

Comprehensive Physiology offers downloadable PowerPoint presentations of figures for non-profit, educational use, provided the content is not modified and full credit is given to the author and publication.

Download a PowerPoint presentation of all images


Figure 1. Figure 1.

Recreational “scuba” (self‐contained underwater breathing apparatus) diver using a standard single‐cylinder and open‐circuit demand valve regulator configuration.

Figure 2. Figure 2.

Diver using a closed‐circuit rebreather. U.S. Navy Experimental Diving Unit photograph.

Figure 3. Figure 3.

Divers in surface supply diving equipment. Each diver has an umbilical supplying gas from the surface and an emergency “bailout” cylinder of gas. Note the oronasal mask visible behind the helmet faceplate of the diver facing the camera. U.S. Navy photograph by Chief Petty Officer Andrew McKaskle.

Figure 4. Figure 4.

Monoplace hyperbaric chamber. The patient is the only occupant. The chamber may be pressurized with air and 100% oxygen delivered via a demand valve and oronasal mask, or the chamber may simply be pressurized with oxygen.

Figure 5. Figure 5.

Multiplace hyperbaric chamber. The chamber is pressurized with air, and delivery of 100% oxygen may be achieved either by the use of a hood sealed around the neck through which oxygen free flows or by a demand valve and oronasal mask.

Figure 6. Figure 6.

Pooled data demonstrating the fall in maximum breathing capacity as ambient pressure increases. The studies were conducted using nonimmersed subjects breathing air.

Reproduced with permission from Camporesi and Bosco 31
Figure 7. Figure 7.

Maximum expiratory flow‐volume curves and isovolume pressure‐flow curves at 75% vital capacity (VC) for a single subject breathing air at 1 and 10 atm abs. In the 1 atm abs condition, a small increase in pleural pressure results in a substantial increase in flow. At 10 atm abs, a small increase in flow requires a much greater increase in pleural pressure, and further increases are limited by the plateau of the isovolume pressure‐flow curve at relatively low flow rates. TLC, total lung capacity.

Reproduced with permission from Wood and Bryan 298
Figure 8. Figure 8.

Schematics to represent static lung loading (or transrespiratory pressure) during immersion. Panel A depicts upright head‐out immersion in which there is a negative static lung load of approximately −20 cmH2O. Panel B depicts an upright scuba diver breathing from a regulator in which there is a similar negative static lung load of approximately −20 cmH2O irrespective of depth. Panel C depicts a closed‐circuit rebreather diver with a back‐mounted counterlung. All of the scenarios depicted in panels (A), (B), and (C) produce a negative static lung load because the airway is in continuity with a gas supply at lower pressure than the hydrostatic pressure acting at the lung centroid. In contrast, Panel D depicts a scuba diver in the head‐down position where there is a positive static lung load because gas supply pressure (measured at the depth of the regulator mouthpiece) is now greater than the hydrostatic pressure acting at the lung centroid. SLL, static lung load.

Reproduced with permission from Lundgren 183
Figure 9. Figure 9.

(A) Dead space (Vd) as a function of tidal volume (Vt) in nonimmersed subjects breathing air at 1 atm abs (“surface”) and during experimental hyperbaric chamber exposures (“depth”) to between 47 and 66 atm abs (inspired gas densities between 12.3 and 17.1 g/liter). (B) The Vd/Vt ratio during increasing levels of exercise at the “surface” and at “depth.” At the surface during exercise at 360 kpm, the ratio decreases by 43% of the resting value in comparison with a 10% decrease at depth.

Reproduced with permission from Salzano et al. 233
Figure 10. Figure 10.

Pooled data from human studies (closed circles) illustrating the effect of gas density (ρ in the formula) on the Vd/Vt ratio as respired gas density increases. The extremely high‐density datapoint (open circle) comes from a liquid‐breathing experiment in dogs.

Reproduced with permission from Moon et al. 200
Figure 11. Figure 11.

Rates of development of pulmonary symptoms and decrements in slow vital capacity in human subjects continuously breathing oxygen at increasing inspired pressures (ATA = atm abs).

Reproduced with permission from Clark et al. 42
Figure 12. Figure 12.

Blood oxygen content for oxygen partial pressures extending into hyperbaric conditions. The right‐angle arrows illustrate oxygen extraction of 5 vol% for air breathing at sea level (left arrow) and for oxygen breathing at 3 atm abs.

Figure 13. Figure 13.

Intercapillary diffusion distances for oxygen depicted using the Krogh cylinder model under conditions of air breathing at 1 atm abs (top left); oxygen breathing at 3 atm abs (top right); and oxygen breathing at 3 atm abs assuming countercurrent flow in adjacent capillaries (bottom). Cylinder boundaries are defined by the calculated distance from capillary to the point where tissue Po2 falls to 12 mmHg. The advantage of a high Pao2, particularly at the arterial end of the capillary is clearly apparent, as is the potential advantage of a countercurrent flow pattern during hyperbaric oxygen breathing.

Reproduced with permission from Saltzman 230
Figure 14. Figure 14.

Ambient pressure (solid line) versus time and the corresponding total gas pressures (ΣPtisj) in two compartments (half‐times of 1 and 5 min) for an air‐breathing dive. Classical decompression is scheduled to keep dissolved gas pressure in k modeled compartments less than or equal to a maximum permissible value, Ptisk ≤ akPamb + bk. For simplicity, the above figure shows only two compartments and sets a = 1 and b = 0 so that the “safe ascent depth”, Pamb = max[(Ptiskb)/a], is equal to max(Ptisk). By convention, decompression stops are taken at increments of 10 fsw (0.3 atm abs) deeper than sea level.

Figure 15. Figure 15.

Simulation of the equilibration of blood with inspired nitrogen with compression breathing air. Simulation is based on the standard, resting 70‐kg man 189 with inspired gas, alveolar gas, pulmonary blood, and the body in series, the latter composed of four parallel compartments representing vessel rich, muscle, fat, and vessel poor tissue groups 74. The simulation uses nitrogen tissue solubility coefficients (α) of 0.015 (blood), 0.015 (lean), and 0.075 (fat) and an estimated lung nitrogen diffusing capacity of 0.15 liters/min/kPa 170,180. A ramp in inspired Pn2 occurs with compression from 101 kPa (sea level) to 404 kPa (4 atm abs) ambient pressure (dashed line). Ninety‐nine percent equilibration of arterial blood with inspired Pn2 (solid line) occurs 2.5 min after reaching depth. Throughout the time course of this simulation, there less than 0.3% difference between arterial blood and alveolar Pn2.

Figure 16. Figure 16.

(A) Washout of nitrogen (sum of dissolved and free gas) from tissue for the cases of all gas remaining dissolved (dotted line) according to Eq. 14, for phase equilibrium between bubble and dissolved gas partial pressures (thin line) according to Eq. 19, and for a single, spherical bubble in a small tissue volume (equivalent to a very high bubble density of 10−6 bubbles/ml, thick line) by using a three‐region model diffusion‐limited bubble growth 245. (B) Dissolved tissue Pn2 (thin line, left axis) and bubble volume (thick line, right axis) for the spherical bubble case illustrated in panel (A).

Figure 17. Figure 17.

The oxygen window. (A) Increasing dissolved oxygen and CO2 concentrations for a liquid at equilibrium with increasing partial pressures of these gases for Ostwald solubility coefficients of 0.024 for oxygen and 0.528 for CO2. (B) Relationship similar to panel A, but with axes reversed. Upper arrow indicates extraction of 2.5 vol% oxygen from the liquid, and lower arrow indicates replacement with 2.5 vol% CO2; the dotted line indicates the resulting oxygen window that is the net difference for the sum of all gas partial pressures in the liquid. Note that extraction of more oxygen, similarly replaced with CO2, would result in a larger oxygen window.

Figure 18. Figure 18.

Change in the sum of all dissolved tissue gas partial pressures (heavy solid line) in response to a compression (dot‐dash) and subsequent change in inspired oxygen fraction (thin solid line, arbitrary scale). The inherent unsaturation (oxygen window) is equal to the vertical distance between the dashed line and the tissue gas partial pressure. Prior to compression, the tissue is inherently unsaturated. For a period following the compression, the total unsaturation (ambient – tissue) exceeds the inherent unsaturation until the alveolar and tissue Pn2 reequilibrate. Increasing the inspired oxygen fraction causes washout of tissue nitrogen and a corresponding increase in the inherent unsaturation. Inher. Unsat., inherent unsaturation.

Figure 19. Figure 19.

The oxygen window increases with increase inspired Po2 because of the nonlinear relationship between blood oxygen content and Po2 due to oxygen‐hemoglobin dissociation (x‐ and y‐axes reversed from the familiar presentation, such as in Figure 12). The two right‐angle arrows indicate the arterial‐venous Po2 differences resulting from 5 vol% oxygen extraction for inspired Po2 of 1.3 or 0.2 atm abs. Since the corresponding increase in Pco2 will be approximately equal in both cases, and assuming equilibrium between tissue and venous blood gases, the difference between the vertical segments of the arrows illustrates the difference in magnitude of the oxygen window for the two cases.

Figure 20. Figure 20.

Whole‐body washout of helium and nitrogen. Y‐axis is the fraction of total expired gas collected. Drawn from data of Behnke and colleagues 17,69.

Figure 21. Figure 21.

Isobaric exchange of helium and nitrogen in a compartment with fivefold difference in time constants (τHe = 7.2, = 36). Panel A is a simulation of a compartment at equilibrium with 90% N2‐10% O2 inspired gas at 10 atm abs ambient pressure and a switch to 90% He‐10% O2 inspired gas at time zero. Dashed and thin lines indicate partial pressures of helium and nitrogen. The thick line indicates the sum of both inert gases and metabolic gases. The compartment is transiently supersaturated, while the sum of gases is above ambient pressure (dotted line). Panel B shows the reverse gas switch.



Figure 1.

Recreational “scuba” (self‐contained underwater breathing apparatus) diver using a standard single‐cylinder and open‐circuit demand valve regulator configuration.



Figure 2.

Diver using a closed‐circuit rebreather. U.S. Navy Experimental Diving Unit photograph.



Figure 3.

Divers in surface supply diving equipment. Each diver has an umbilical supplying gas from the surface and an emergency “bailout” cylinder of gas. Note the oronasal mask visible behind the helmet faceplate of the diver facing the camera. U.S. Navy photograph by Chief Petty Officer Andrew McKaskle.



Figure 4.

Monoplace hyperbaric chamber. The patient is the only occupant. The chamber may be pressurized with air and 100% oxygen delivered via a demand valve and oronasal mask, or the chamber may simply be pressurized with oxygen.



Figure 5.

Multiplace hyperbaric chamber. The chamber is pressurized with air, and delivery of 100% oxygen may be achieved either by the use of a hood sealed around the neck through which oxygen free flows or by a demand valve and oronasal mask.



Figure 6.

Pooled data demonstrating the fall in maximum breathing capacity as ambient pressure increases. The studies were conducted using nonimmersed subjects breathing air.

Reproduced with permission from Camporesi and Bosco 31


Figure 7.

Maximum expiratory flow‐volume curves and isovolume pressure‐flow curves at 75% vital capacity (VC) for a single subject breathing air at 1 and 10 atm abs. In the 1 atm abs condition, a small increase in pleural pressure results in a substantial increase in flow. At 10 atm abs, a small increase in flow requires a much greater increase in pleural pressure, and further increases are limited by the plateau of the isovolume pressure‐flow curve at relatively low flow rates. TLC, total lung capacity.

Reproduced with permission from Wood and Bryan 298


Figure 8.

Schematics to represent static lung loading (or transrespiratory pressure) during immersion. Panel A depicts upright head‐out immersion in which there is a negative static lung load of approximately −20 cmH2O. Panel B depicts an upright scuba diver breathing from a regulator in which there is a similar negative static lung load of approximately −20 cmH2O irrespective of depth. Panel C depicts a closed‐circuit rebreather diver with a back‐mounted counterlung. All of the scenarios depicted in panels (A), (B), and (C) produce a negative static lung load because the airway is in continuity with a gas supply at lower pressure than the hydrostatic pressure acting at the lung centroid. In contrast, Panel D depicts a scuba diver in the head‐down position where there is a positive static lung load because gas supply pressure (measured at the depth of the regulator mouthpiece) is now greater than the hydrostatic pressure acting at the lung centroid. SLL, static lung load.

Reproduced with permission from Lundgren 183


Figure 9.

(A) Dead space (Vd) as a function of tidal volume (Vt) in nonimmersed subjects breathing air at 1 atm abs (“surface”) and during experimental hyperbaric chamber exposures (“depth”) to between 47 and 66 atm abs (inspired gas densities between 12.3 and 17.1 g/liter). (B) The Vd/Vt ratio during increasing levels of exercise at the “surface” and at “depth.” At the surface during exercise at 360 kpm, the ratio decreases by 43% of the resting value in comparison with a 10% decrease at depth.

Reproduced with permission from Salzano et al. 233


Figure 10.

Pooled data from human studies (closed circles) illustrating the effect of gas density (ρ in the formula) on the Vd/Vt ratio as respired gas density increases. The extremely high‐density datapoint (open circle) comes from a liquid‐breathing experiment in dogs.

Reproduced with permission from Moon et al. 200


Figure 11.

Rates of development of pulmonary symptoms and decrements in slow vital capacity in human subjects continuously breathing oxygen at increasing inspired pressures (ATA = atm abs).

Reproduced with permission from Clark et al. 42


Figure 12.

Blood oxygen content for oxygen partial pressures extending into hyperbaric conditions. The right‐angle arrows illustrate oxygen extraction of 5 vol% for air breathing at sea level (left arrow) and for oxygen breathing at 3 atm abs.



Figure 13.

Intercapillary diffusion distances for oxygen depicted using the Krogh cylinder model under conditions of air breathing at 1 atm abs (top left); oxygen breathing at 3 atm abs (top right); and oxygen breathing at 3 atm abs assuming countercurrent flow in adjacent capillaries (bottom). Cylinder boundaries are defined by the calculated distance from capillary to the point where tissue Po2 falls to 12 mmHg. The advantage of a high Pao2, particularly at the arterial end of the capillary is clearly apparent, as is the potential advantage of a countercurrent flow pattern during hyperbaric oxygen breathing.

Reproduced with permission from Saltzman 230


Figure 14.

Ambient pressure (solid line) versus time and the corresponding total gas pressures (ΣPtisj) in two compartments (half‐times of 1 and 5 min) for an air‐breathing dive. Classical decompression is scheduled to keep dissolved gas pressure in k modeled compartments less than or equal to a maximum permissible value, Ptisk ≤ akPamb + bk. For simplicity, the above figure shows only two compartments and sets a = 1 and b = 0 so that the “safe ascent depth”, Pamb = max[(Ptiskb)/a], is equal to max(Ptisk). By convention, decompression stops are taken at increments of 10 fsw (0.3 atm abs) deeper than sea level.



Figure 15.

Simulation of the equilibration of blood with inspired nitrogen with compression breathing air. Simulation is based on the standard, resting 70‐kg man 189 with inspired gas, alveolar gas, pulmonary blood, and the body in series, the latter composed of four parallel compartments representing vessel rich, muscle, fat, and vessel poor tissue groups 74. The simulation uses nitrogen tissue solubility coefficients (α) of 0.015 (blood), 0.015 (lean), and 0.075 (fat) and an estimated lung nitrogen diffusing capacity of 0.15 liters/min/kPa 170,180. A ramp in inspired Pn2 occurs with compression from 101 kPa (sea level) to 404 kPa (4 atm abs) ambient pressure (dashed line). Ninety‐nine percent equilibration of arterial blood with inspired Pn2 (solid line) occurs 2.5 min after reaching depth. Throughout the time course of this simulation, there less than 0.3% difference between arterial blood and alveolar Pn2.



Figure 16.

(A) Washout of nitrogen (sum of dissolved and free gas) from tissue for the cases of all gas remaining dissolved (dotted line) according to Eq. 14, for phase equilibrium between bubble and dissolved gas partial pressures (thin line) according to Eq. 19, and for a single, spherical bubble in a small tissue volume (equivalent to a very high bubble density of 10−6 bubbles/ml, thick line) by using a three‐region model diffusion‐limited bubble growth 245. (B) Dissolved tissue Pn2 (thin line, left axis) and bubble volume (thick line, right axis) for the spherical bubble case illustrated in panel (A).



Figure 17.

The oxygen window. (A) Increasing dissolved oxygen and CO2 concentrations for a liquid at equilibrium with increasing partial pressures of these gases for Ostwald solubility coefficients of 0.024 for oxygen and 0.528 for CO2. (B) Relationship similar to panel A, but with axes reversed. Upper arrow indicates extraction of 2.5 vol% oxygen from the liquid, and lower arrow indicates replacement with 2.5 vol% CO2; the dotted line indicates the resulting oxygen window that is the net difference for the sum of all gas partial pressures in the liquid. Note that extraction of more oxygen, similarly replaced with CO2, would result in a larger oxygen window.



Figure 18.

Change in the sum of all dissolved tissue gas partial pressures (heavy solid line) in response to a compression (dot‐dash) and subsequent change in inspired oxygen fraction (thin solid line, arbitrary scale). The inherent unsaturation (oxygen window) is equal to the vertical distance between the dashed line and the tissue gas partial pressure. Prior to compression, the tissue is inherently unsaturated. For a period following the compression, the total unsaturation (ambient – tissue) exceeds the inherent unsaturation until the alveolar and tissue Pn2 reequilibrate. Increasing the inspired oxygen fraction causes washout of tissue nitrogen and a corresponding increase in the inherent unsaturation. Inher. Unsat., inherent unsaturation.



Figure 19.

The oxygen window increases with increase inspired Po2 because of the nonlinear relationship between blood oxygen content and Po2 due to oxygen‐hemoglobin dissociation (x‐ and y‐axes reversed from the familiar presentation, such as in Figure 12). The two right‐angle arrows indicate the arterial‐venous Po2 differences resulting from 5 vol% oxygen extraction for inspired Po2 of 1.3 or 0.2 atm abs. Since the corresponding increase in Pco2 will be approximately equal in both cases, and assuming equilibrium between tissue and venous blood gases, the difference between the vertical segments of the arrows illustrates the difference in magnitude of the oxygen window for the two cases.



Figure 20.

Whole‐body washout of helium and nitrogen. Y‐axis is the fraction of total expired gas collected. Drawn from data of Behnke and colleagues 17,69.



Figure 21.

Isobaric exchange of helium and nitrogen in a compartment with fivefold difference in time constants (τHe = 7.2, = 36). Panel A is a simulation of a compartment at equilibrium with 90% N2‐10% O2 inspired gas at 10 atm abs ambient pressure and a switch to 90% He‐10% O2 inspired gas at time zero. Dashed and thin lines indicate partial pressures of helium and nitrogen. The thick line indicates the sum of both inert gases and metabolic gases. The compartment is transiently supersaturated, while the sum of gases is above ambient pressure (dotted line). Panel B shows the reverse gas switch.

References
 1. Ackles KN, Holness DE, Scott CA. Measurement of uptake and elimination of nitrogen in tissue, in vivo. In: Lambertsen CJ, editor. Underwater Physiology V, Proceedings of the 5th Symposium on Underwater Physiology. Bethesda, MD: Federation of American Societies for Experimental Biology, 1976, p. 349‐354.
 2. Agostoni E, Gurtner G, Torri G, Rahn H. Respiratory mechanics during submersion and negative pressure breathing. J Appl Physiol 21: 251‐258, 1966.
 3. Albin GW, Weathersby PK. Statistically Based Decompression Tables VI. Repeat Dives on Oxygen/Nitrogen Mixes (Technical Report 91‐84). Bethesda, MD: Naval Medical Research Institute, 1991.
 4. Anderson D, George J, Lundgren CEG. Moderate hypercapnia: Cardiovascular function and nitrogen elimination. Undersea Hyperb Med 20: 225‐232, 1993.
 5. Anthonisen NR, Utz G, Kryger MH, Urbanetti JS. Exercise tolerance at 4 and 6 atm abs. Undersea Biomed Res 3: 95‐102, 1976.
 6. Arieli R. Cyclic perfusion of the lung by dense gas breathing may reduce the (A‐a)DO2. J Basic Clin Physiol Pharmacol 3: 207‐221, 1992.
 7. Arieli R, Farhi LE. Gas exchange in tidally ventilated and non‐steadily perfused lung model. Respir Physiol 60: 295‐309, 1985.
 8. Aukland K, Akre S, Leraand S. Arteriovenous counter‐current exchange of hydrogen gas in skeletal muscle. Scand J Clin Lab Invest Suppl 99: 72‐75, 1967.
 9. Babb TG, Viggiano R, Hurley B, Staarts B, Rodarte JR. Effect of mild to moderate airflow limitation on exercise capacity. J Appl Physiol 70: 223‐230, 1991.
 10. Balldin UI. Effects of ambient temperature and body position on tissue nitrogen elimination in man. Aerosp Med 44: 365‐370, 1973.
 11. Balldin UI, Lundgren CE. Effects of immersion with the head above water on tissue nitrogen elimination in man. Aerosp Med 43: 1101‐1108, 1972.
 12. Barnett TB, Rasmussen B. Ventilatory responses to hypoxia and hypercapnia with external airway resistance. Acta Physiol Scand 80: 538‐551, 1970.
 13. Becker HF, Polo O, McNamara SG, Bethon‐Jones M, Sullivan CE. Effect of different levels of hyperoxia on breathing in healthy subjects. J Appl Physiol 81: 1683‐1690, 1996.
 14. Behnke AR. The isobaric (oxygen window) principle of decompression. In: The New Thrust Seaward, Transactions of the Third Annual Conference of the Marine Technology Society. Washington, DC: Marine Technology Society, 1967.
 15. Behnke AR, Thomson RM, Motley EP. The psychological effects from breathing at 4 atmospheres pressure. Am J Physiol 112: 554‐558, 1935.
 16. Behnke AR, Thomson RM, Shaw LA. The rate of elimination of dissolved nitrogen in man in relation to the fat and water content of the body. Am J Physiol 114: 137‐146, 1935.
 17. Behnke AR, Willmon TL. Gaseous nitrogen and helium elimination from the body during rest and exercise. Am J Physiol 131: 619‐626, 1940.
 18. Bennett M, Lehm J, Barr P. Medical support for the Sydney Airport Link Tunnel Project. South Pacific Underwater Med Soc J 32: 90‐96, 2002.
 19. Bennett PB, Vann RD, Roby J, Youngblood D. Theory and development of subsaturation decompression procedures for depths in excess of 400 feet. In: Shilling CW, Beckett MW, editors. Underwater Physiology VI, Proceedings of the 6th Symposium on Underwater Physiology. Bethesda, MD: Federation of American Societies for Experimental Biology, 1978, p. 367‐381.
 20. Blake RB, Morin RA. Pressure increases oxygen affinity of whole blood and erythrocyte suspensions. J Appl Physiol 61: 486‐94, 1986.
 21. Boerema I, Mayne NG, Brummelkamp WK, Bouma S, Mensch MH, Kamermans F, Stern Hanf M, van Aalderen W. Life without blood: A study of the influence of high atmospheric pressure and hypothermia on dilution of the blood. J Cardiovasc Surg 1: 133‐146, 1960.
 22. Boycott AE, Damant GCC, Haldane JS. The prevention of compressed‐air illness. J Hygiene (Lond) 8: 342‐443, 1908.
 23. Brodersen P, Sejrsen P, Lassen NA. Diffusion bypass of xenon in brain circulation. Circ Res 32: 363‐369, 1973.
 24. Bühlmann AA. Die Berechnung der risikoarmen Dekompression. Schweiz Med Wochenschr 118: 185‐197, 1988.
 25. Bühlmann AA, Gehring H. Inner ear disorders resulting from inadequate decompression “vertigo bends.” In: Lambertsen CJ, editor. Underwater Physiology V, Proceedings of the 5th Symposium on Underwater Physiology. Bethesda, MD: Federation of American Societies for Experimental Biology, 1976, p. 341‐347.
 26. Burkard ME, Van Liew HD. Simulation of exchanges of multiple gases in bubbles in the body. Respir Physiol 95: 131‐145, 1994.
 27. Cain CC, Otis AB. Some physiological effects resulting from added resistance to respiration. J Aviat Med 20: 149‐160, 1949.
 28. Cain SM. Gas exchange in hypoxia, apnea, and hyperoxia. In: Farhi LE, Tenney SM, editors. Handbook of Physiology. Bethesda, MD: American Physiological Society, 1987, sect. 3, vol. IV, chapt. 19, p. 403‐420.
 29. Caldwell PRB, Lee WL Jr, Schildkraut HS, Archibald ER. Changes in lung volume, diffusing capacity, and blood gases in men breathing oxygen. J Appl Physiol 21: 1477‐1483, 1966.
 30. Campbell JA, Hill L. Studies in the saturation of the tissues with gaseous nitrogen, part I: Rate of saturation of goat's bone marrow in vivo with nitrogen during exposure to increased atmospheric pressure. Q J Exp Biol 23: 197‐210, 1933.
 31. Camporesi EM, Bosco G. Ventilation, gas exchange, and exercise under pressure. In: Brubakk AO, Neuman TS, editors. Bennett and Elliott's Physiology and Medicine of Diving (5th ed). Edinburgh, UK: Saunders, 2003, p. 77‐114.
 32. Cerretelli P, Di Prampero PE. Gas exchange in exercise. In: Farhi LE, Tenney SM, editors. Handbook of Physiology. Bethesda, MD: American Physiological Society 1987, sect. 3, vol. IV, chapt. 16, p. 297‐340.
 33. Cerretelli P, Sikand SS, Farhi LE. Effect of increased airway resistance on ventilation and gas exchange during exercise. J Appl Physiol 27: 597‐600, 1969.
 34. Chavko M, Nemoto EM, Melick JA. Regional lipid composition in the rat brain. Mol Chem Neuropathol 18: 123‐131, 1993.
 35. Cherniak RM, Snidal DP. The effect of obstruction to breathing on the ventilatory response to CO2.. J Clin Invest 35: 1286‐1290, 1956.
 36. Cherry AD, Forkner IF, Frederick HJ, Natoli MJ, Schinazi EA, Longphre JP, Conard JL, White WD, Freiberger JJ, Stolp BW, Pollock NW, Doar PO, Boso AE, Alford EL, Walker AJ, Ma AC, Rhodes MA, Moon RE. Predictors of increased Paco2 during immersed prone exercise at 4.7 atm abs. J Appl Physiol 106: 316‐325, 2009.
 37. Christopherson SK, Hlastala MP. Pulmonary gas exchange during altered density gas breathing. J Appl Physiol 52: 221‐225, 1982.
 38. Clark JM. Oxygen toxicity. In: Neuman TS, Thom SR, editors. The Physiology and Medicine of Hyperbaric Oxygen Therapy. Philadelphia: Saunders, 2008, p. 527‐564.
 39. Clark JM, Gelfand R, Lambertsen CJ, Stevens WC, Beck G Jr, Fisher DG. Human tolerances and physiological responses to exercise while breathing oxygen at 2.0 atm abs. Aviat Space Environ Med 66: 336‐345, 1995.
 40. Clark JM, Jackson RM, Lambertsen CJ, Gelfand R, Hiller WDB, Unger M. Pulmonary function in men after oxygen breathing at 3.0 atm abs for 3.5 hours. J Appl Physiol 71: 878‐885, 1991.
 41. Clark JM, Lambertsen CJ. Alveolar‐arterial O2 differences in man at 0.2, 1.0, 2.0, and 3.5 atm abs inspired Po2. J Appl Physiol 30: 753‐763, 1971.
 42. Clark JM, Lambertsen CJ, Gelfand R, Flores ND, Pisarello JB, Rossman MD, Elias JA. Effects of prolonged oxygen exposure at 1.5, 2.0, or 2.5 atm abs on pulmonary function in men (predictive studies V). J Appl Physiol 86: 243‐59, 1999.
 43. Clark JM, Thom SR. Oxygen under pressure. In: Brubakk AO, Neuman TS, editors. Bennett and Elliott's Physiology and Medicine of Diving (5th ed). Edinburgh, UK: Saunders, 2003, p. 358‐417.
 44. Clarke JR. Underwater breathing apparatus. In: Lundgren CEG, Miller JN, editors. The Lung at Depth. New York: Marcel Decker, 1999, p. 429‐527.
 45. Clarke JR, Flook V. Respiratory function at depth. In: Lundgren CEG, Miller JN, editors. The Lung at Depth. New York: Marcel Decker, 1999, p. 1‐71.
 46. Clarke JR, Jaeger MJ, Zumrick JL, O'Bryan R, Spaur WH. Respiratory resistance from 1 to 46 atm abs measured with the interrupter technique. J Appl Physiol 52: 549‐555, 1982.
 47. Cohen R, Bell WH, Saltzman HA, Kylstra JA. Alveolar‐arterial oxygen pressure difference in man immersed up to the neck in water. J Appl Physiol 30: 720‐723, 1971.
 48. Cropp GJA. Effect of high intra‐alveolar O2 tensions on pulmonary circulation in perfused lungs of dogs. Am J Physiol 208: 130‐138, 1965.
 49. Dahan A, DeGoede J, Berkenbosch A, Olievier ICW. The Influence of oxygen on the ventilatory response to carbon dioxide in man. J Physiol 428: 485‐499, 1990.
 50. Dahlback GO, Jonsson E, Liner MH. Influence of hydrostatic compression of the chest and intrathoracic blood pooing on static lung mechanics during head‐out immersion. Undersea Biomed Res 5: 71‐85, 1978.
 51. D'Aoust BG, Smith KH, Swanson HT. Decompression‐induced decrease in nitrogen elimination rate in awake dogs. J Appl Physiol 41: 348‐355, 1976.
 52. D'Aoust BG, Smith KH, Swanson HT, White R, Stayton L, Moore J. Prolonged bubble production by transient isobaric counter‐equilibration of helium against nitrogen. Undersea Biomed Res 6: 109‐125, 1979.
 53. D'Aoust BG, Swanson HT, White R, Dunford R, Mahoney J. Central venous bubbles and mixed venous nitrogen in goats following decompression. J Appl Physiol 51: 1238‐1244, 1981.
 54. Dawson SV, Elliott EA. Wave‐speed limitation on expiratory flow—a unifying concept. J Appl Physiol 43: 498‐515, 1977.
 55. Demchenko IT, Boso AE, Bennett PB, Whorton AR, Piantadosi CA. Hyperbaric oxygen reduced cerebral blood flow by inactivating nitric oxide. Nitric Oxide 4: 597‐608, 2000.
 56. Demchenko IT, Oury TD, Crapo JD, Piantadosi CA. Regulation of brain's vascular responses to oxygen. Circ Res 91: 1031‐1037, 2002.
 57. Demedts M, Anthonisen NR. Effects of increased external airway resistance during steady‐state exercise. J Appl Physiol 35: 361‐366, 1973.
 58. Deming WE, Shupe LE. Some physical properties of compressed gases I. Nitrogen. Phy Rev 37: 638‐654, 1931.
 59. Derion T, Guy HJB, Tsukimoto K, Schaffartzik W, Prediletto R, Poole DC, Knight DR, Wagner PD. Ventilation‐perfusion relationships in the lung during head‐out immersion. J Appl Physiol 72: 64‐72, 1992.
 60. Derion T, Guy HJB. Effects of age on closing volume during head‐out water immersion. Respir Physiol 95: 273‐280, 1994.
 61. Doolette DJ, Mitchell SJ. The physiological kinetics of nitrogen and the prevention of decompression illness. Clin Pharmacokinet 40: 1‐14, 2001.
 62. Doolette DJ, Mitchell SJ. A biophysical basis for inner ear decompression sickness. J Appl Physiol 94: 2145‐2150, 2003.
 63. Doolette DJ, Upton RN, Grant C. Diffusion limited, but not perfusion limited, compartmental models describe cerebral nitrous oxide kinetics at both high and low cerebral blood flows. J Pharmacokinet Biopharm 26: 649‐672, 1998.
 64. Doolette DJ, Upton RN, Grant C. Isobaric exchange of helium and nitrogen in the brain at high and low blood flow [abstract]. Undersea Hyperb Med 31: 340, 2004.
 65. Doolette DJ, Upton RN, Grant C. Countercurrent compartmental models describe hind limb skeletal muscle helium kinetics at resting and low blood flows in sheep. Acta Physiol Scand 185: 109‐121, 2005.
 66. Doolette DJ, Upton RN, Grant C. Isobaric exchange of helium and nitrogen in skeletal muscle at resting and low blood flow [abstract]. Undersea Hyperb Med 32: 307, 2005.
 67. Doolette DJ, Upton RN, Grant C. Perfusion‐diffusion compartmental models describe cerebral helium kinetics at high and low cerebral blood flows in sheep. J Physiol (Lond) 563: 529‐539, 2005.
 68. Doolette DJ, Upton RN, Zheng D. Diffusion‐limited tissue equilibration and arteriovenous diffusion shunt describe skeletal muscle nitrous oxide kinetics at high and low blood flows in sheep. Acta Physiol Scand 172: 167‐177, 2001.
 69. Duffner GJ, Snider HH. Effects of Exposing Men to Compressed Air and Helium‐Oxygen Mixtures for 12 Hours at Pressures of 2–2.6 Atmospheres (Technical Report 1‐59). Panama City, FL: Navy Experimental Diving Unit, 1958.
 70. D'Urzo AD, Chapman KR, Rebuck AS. Effect of inspiratory resistive loading on control of ventilation during progressive exercise. J Appl Physiol 62: 134‐140, 1987.
 71. Dwyer J, Saltzman HA, O'Bryan R. Maximum physical‐work capacity of man at 43.4 atm abs. Undersea Biomed Res 4: 359‐372, 1977.
 72. Dwyer JV. Calculation of Repetitive Diving Decompression Tables (Research Report 1‐57). Washington, DC: Navy Experimental Diving Unit, 1956.
 73. Eckenhoff RG, Olstad CS, Carrod G. Human dose‐response relationship for decompression and endogenous bubble formation. J Appl Physiol 69: 914‐918, 1990.
 74. Eger EI. A mathematical model of uptake and distribution. In: Papper EM, Kitz RJ, editors. Uptake and Distribution of Anesthetic Agents. New York: McGraw‐Hill, 1962, p. 72‐87.
 75. Ehler WJ, Marx RE, Peleo MJ. Oxygen as a drug: A dose response curve for radiation necrosis. Undersea Hyper Med 20 (Suppl): 94‐95, 1993.
 76. Eldridge F, Davis JM. Effect of mechanical factors on respiratory work and ventilatory responses to CO2. J Appl Physiol 14: 721‐726, 1959.
 77. Elliott DH. Loss of consciousness underwater. In: Bennett PB, Moon RE, editors. Diving Accident Management. Proceedings of the 41st Undersea and Hyperbaric Medical Society Workshop. Bethesda, MD: Undersea and Hyperbaric Medical Society, 1990, p. 301‐310.
 78. Epstein PS, Plesset MS. On the stability of gas bubbles in liquid‐gas solutions. J Chem Phys 18: 1505‐1509, 1950.
 79. Evans A, Walder DN. Significance of gas micronuclei in the aetiology of decompression sickness. Nature 222: 251‐252, 1969.
 80. Fagraeus L, Hesser CM. Ventilatory response to CO2 in hyperbaric environments. Acta Physiol Scand 80: 19A‐20A, 1970.
 81. Fagraeus L, Hesser CM, Linnarsson D. Cardiorespiratory responses to graded exercise at increased ambient air pressure. Acta Physiol Scand 91: 259‐74, 1974.
 82. Fagraeus L, Linnarsson D. Maximal voluntary and exercise ventilation at high ambient air pressures. Forsvarmedicin 9: 275‐278, 1973.
 83. Farmer JC Jr. Diving injuries to the inner ear. Ann Otol Rhinol Laryngol Suppl 86: 1‐20, 1977.
 84. Feldmeier JJ. Hyperbaric oxygen therapy for delayed radiation injuries. In: Neuman TS, Thom SR, editors. The Physiology and Medicine of Hyperbaric Oxygen Therapy. Philadelphia: Saunders, 2008, p. 231‐256.
 85. Flook V, Kelman GR. Submaximal exercise with increased inspiratory resistance to breathing. J Appl Physiol 35: 379‐384, 1973.
 86. Florio JT, Morrison JB, Butt WS. Breathing pattern and ventilatory response to carbon dioxide in divers. J Appl Physiol 46: 1076‐1080, 1979.
 87. Flynn ET, Saltzman HA, Summitt JK. Effects of head‐out immersion at 19.18 atm abs on pulmonary gas exchange in man. J Appl Physiol 33: 113‐119, 1972.
 88. Forkner IF, Piantadosi CA, Scafetta N, Moon RE. Hyperoxia‐induced tissue hypoxia: A danger? Anesthesiology 106: 1051‐1055, 2007.
 89. Forster RE. Diffusion of gases across the alveolar membrane. In: Farhi LE, Tenney SM, editors. Handbook of Physiology. Bethesda, MD: American Physiological Society, 1987, sect. 3, vol. IV, chapt. 5, p. 71‐88.
 90. Fothergill DM, Hedges D, Morrison JM. Effects of CO2 and N2 partial pressures on cognitive and psychomotor performance. Undersea Biomed Res 18: 1‐19, 1991.
 91. Fothergill DM, Joye DD, Carlson NA. Diver respiratory response to a tunable closed‐circuit breathing apparatus. Undersea Hyperb Med 24: 91‐105, 1997.
 92. Francis TJR, Mitchell SJ. Manifestations of decompression disorders. In: Brubakk AO, Neuman TS, editors. Bennett and Elliott's Physiology and Medicine of Diving (5th ed). Edinburgh: Saunders, 2003, p. 578‐599.
 93. Francis TJR, Mitchell SJ. The pathophysiology of decompression sickness. In: Brubakk AO, Neuman TS, editors. Bennett and Elliott's Physiology and Medicine of Diving (5th ed). Edinburgh: Saunders, 2003, p. 530‐556.
 94. Gardette B, Fructus X, Delauze HG. First human hydrogen saturation dives at 450 msw: Hydra V. In: Bove AA, Bachrach AJ, Greenbaum LJ, editors. Proceedings of the 9th International Symposium on Underwater and Hyperbaric Physiology. Bethesda, MD: Undersea and Hyperbaric Medical Society, 1987, p. 375‐389.
 95. Gelfand R, Lambertsen CJ. Dynamic respiratory response to abrupt change of inspired CO2 at normal and high PO2. J Appl Physiol 35: 903‐913, 1973.
 96. Gelfand R, Lambertsen CJ, Peterson RE. Human respiratory control at high ambient pressures and inspired gas densities. J Appl Physiol 48: 528‐539, 1980.
 97. Gelfand R, Lambertsen CJ, Strauss R, Clark JM, Puglia CD. Human respiration at rest in rapid compression and at high pressures and gas densities. J Appl Physiol 54: 290‐303, 1983.
 98. Gernhardt ML. Development and Evaluation of a Decompression Stress Index Based on Tissue Bubble Dynamics (Dissertation). Philadelphia: University of Pennsylvania, 1991.
 99. Gerth WA. Effects of dissolved electrolytes on the solubility and partial molar volumeof helium in water from 50 to 400 atmospheres at 25°C. J Sol Chem 12: 655‐669, 1983.
 100. Gerth WA. Applicability of Henry's law to hydrogen, helium, and nitrogen solubilities and water and olive oil at 37°C and pressures up to 300 atmospheres. Arch Biochem Biophys 241: 187‐199, 1985.
 101. Gerth WA, Doolette DJ, Gault KA. Deep stops and their efficacy in decompression: U.S. Navy research. In: Bennett PB, Wienke BR, Mitchell SJ, editors. Decompression and the Deep Stop, Proceedings of the Undersea and Hyperbaric Medical Society Workshop. Durhham, NC: Undersea and Hyperbaric Medical Society, 2009, p. 165‐185.
 102. Gerth WA, Vann RD. Development of Iso‐DCS Risk Air and Nitrox Decompression Tables Using Statistical Bubble Dynamics Models (Final Report) Rockville, MD: National Oceanic and Atmospheric Administration, Office of Undersea Research, 1996.
 103. Gerth WA, Vann RD. Probabilistic gas and bubble dynamics models of decompression sickness occurrence in air and N2‐O2 diving. Undersea Hyperb Med 24: 275‐292, 1997.
 104. Gledhill N, Froese AB, Buick FJ, Bryan AC. VA/Q inhomogeneity and AaDO2 in man during exercise: Effect of SF6 breathing. J Appl Physiol 45: 512‐515, 1978.
 105. Goldman S. A new class of biophysical models for predicting the probability of decompression sickness in scuba diving. J Appl Physiol 103: 484‐493, 2007.
 106. Graves DJ, Idicula J, Lambertsen CJ, Quin JA. Bubble formation resulting from counterdiffusion supersaturation: A possible explanation for isobaric inert gas ‘urticaria’ and vertigo. Phys Med Biol 18: 256‐264, 1973.
 107. Greenbaum R, Kelman GR, Nunn JF, Ptys‐Roberts C. Arterial oxygen tensions. Anesthesiology 70: 869‐870, 1966.
 108. Grindlay J, Mitchell SJ. Isolated pulmonary oedema associated with scuba diving. Emerg Med 11: 272‐6, 1999.
 109. Groom AC, Morin R, Farhi LE. Determination of dissolved N2 in blood and investigation of N2 washout from the body. J Appl Physiol 23: 706‐712, 1967.
 110. Hales JRS. Effect of exposure to hot environments on total and regional blood flow in the brain and spinal cord of the sheep. Pflügers Arch 344: 327‐337, 1973.
 111. Hamilton RW, Adams GM, Harvey CA, Knight DR. SHAD‐NISAT: A Composite Study of Shallow Saturation Diving 985. Groton, CT: Naval Submarine Medical Research laboratory, 1982.
 112. Hamilton RW, Thalmann ED. Decompression practice. In: Brubakk AO, Neuman TS, editors. Bennett and Elliott's Physiology and Medicine of Diving (5th ed). Edinburgh, UK: Saunders, 2003.
 113. Hampson NB. Hyperbaric Oxygen Therapy: 1999 Committee Report. Kensington, MD: Undersea & Hyperbaric Medical Society, 1999.
 114. Hampson NB, Dunford RG. Pulmonary edema of scuba divers. Undersea Hyper Med 24: 29‐33, 1997.
 115. Harabin AL, Homer LD, Bradley ME. Pulmonary oxygen toxicity in awake dogs: Metabolic and physiological effects. J Appl Physiol 57: 1480‐1488, 1984.
 116. Harvey EN, Whiteley AH, McElroy WD, Pease DC, Barnes DK. Bubble formation in animals, part II: Gas nuclei and their distribution in blood and tissues. J Cell Comp Physiol 24: 23‐34, 1944.
 117. Hashimoto A, Daskalovic I, Reddan WG, Lanphier EH. Detection and modification of CO2 retention in divers. Undersea Biomed Res 8 (1, Suppl): 47, 1981
 118. Hawkins JA, Shilling CW, Hansen RA. A suggested change in calculating decompression tables for diving. US Nav Med Bull 33: 327‐338, 1935.
 119. Hays JR, Hart BL, Weathersby PK, Survanshi SS, Homer LD, Flynn ET. Statistically Based Decompression Tables IV. Extension to Air and N2‐O2 Decompression (Technical Report 86‐51). Bethesda, MD: Naval Medical Research Institute, 1986.
 120. Hemmingsen BB, Steinberg NK, Hemmingsen EA. Intracellular gas supersaturation tolerances of erythrocytes and resealed ghosts. Biophys J 47: 491‐496, 1985.
 121. Hemmingsen EA. Cavitation in gas‐supersatured solutions. J Appl Phys 64: 213‐218, 1975.
 122. Hemmingsen EA. Effects of surfactants and electrolytes on the nucleation of bubbles in gas‐supersaturated solutions. Z Naturforsch 33a: 164‐171, 1978.
 123. Hemmingsen EA, Hemmingsen BB. Lack of intracellular bubble formation in microorganisms at very high gas supersaturations. J Appl Physiol 47: 1270‐1277, 1979.
 124. Hempleman HV. British decompression theory and practice. In: Bennett PB, Elliott DH, editors. The Physiology and Medicine of Diving and Compressed Air Work. London: Ballière Tindall & Cassell, 1969, p. 291‐318.
 125. Hennessy TR. The equivalent bulk‐diffusion model of the pneumatic decompression computer. Med Biol Eng 11: 135‐137, 1973.
 126. Hesser CM, Adolfson J, Fagraeus L. Role of CO2 in compressed air narcosis. Aerospace Med 42: 163‐168, 1971.
 127. Hesser CM, Lind F, Linnarsson D. Significance of airway resistance for the pattern of breathing and lung volumes in exercising humans. J Appl Physiol 68: 1875‐1882, 1990.
 128. Hesser CM, Linnarsson D, Fagraeus L. Pulmonary mechanics and work of breathing at maximal ventilation and raised air pressure. J Appl Physiol 50: 747‐753, 1981.
 129. Hickey DD, Norfleet WT, Pasche AJ, Lundgren CEG. Respiratory function in the upright working diver at 6.8 atm abs (190fsw). Undersea Biomed Res 14: 241‐262, 1987.
 130. Higuchi H, Adachi Y, Arimura S, Kanno M, Satoh T. The carbon dioxide absorption capacity of Amsorb is half that of soda lime. Anesth Analg 93: 221‐225, 2001.
 131. Hill L, Phillips AE. Deep sea diving. J R Nav Med Serv 18: 157‐183, 1932.
 132. Hills BA. A Thermodynamic and Kinetic Approach to Decompression Sickness (Dissertation). Adelaide, South Australia, Australia: The University of Adelaide, 1966.
 133. Hills BA. Decompression Sickness: The Biophysical Basis of Prevention and Treatment. Chichester, UK: John Wiley & Sons, 1977.
 134. Hills BA. Effect of decompression per se on nitrogen elimination. J Appl Physiol 45: 916‐921, 1978.
 135. Hills BA, LeMessurier DH. Unsaturation in living tissue relative to the pressure and composition of inhaled gas and its significance in decompression theory. Clin Sci 36: 185‐195, 1969.
 136. Himm JF, Homer LD. A model of extravascular bubble evolution: Effect of changes in breathing gas composition. J Appl Physiol 87: 1521‐1531, 1999.
 137. Hlastala MP. Diffusing‐capacity heterogeneity. In: Farhi LE, Tenney SM, editors. Handbook of Physiology. Bethesda, MD: American Physiological Society, 1987, sect. 3, vol. IV, chapt. 12, p. 217‐232.
 138. Hof IM, West VP, Younes M. Steady‐state response of normal subjects to inspiratory resistive load. J Appl Physiol 60: 1471‐1481, 1986.
 139. Homer LD, Weathersby PK, Survanshi S. How countercurrent blood flow and uneven perfusion affect motion of inert gas. J Appl Physiol 69: 162‐170, 1990.
 140. Hong SK, Cerretelli P, Cruz JC, Rahn H. Mechanics of respiration during submersion in water. J Appl Physiol 27: 535‐538, 1969.
 141. Hrncir E, Rosin J. Surface tension of blood. Physiol Res 46: 319‐321, 1997.
 142. Hyldegaard O, Jensen T. Effect of heliox, oxygen and air breathing on helium bubbles after heliox diving. Undersea Hyperb Med 34: 107‐122, 2007.
 143. Hyldegaard O, Madsen J. Influence of heliox, oxygen, and N2O‐O2 breathing on N2 bubbles in adipose tissue. Undersea Biomed Res 16: 185‐193, 1989.
 144. Hyldegaard O, Madsen J. Effect of air, heliox, and oxygen breathing on air bubbles in aqueous tissues in the rat. Undersea Hyperb Med 21: 413‐424, 1994.
 145. Hyldegaard O, Moller M, Madsen J. Effect of He‐O2, O2, and N2O‐O2 breathing on injected bubbles in spinal white matter. Undersea Biomed Res 18: 361‐371, 1991.
 146. Iscoe S, Fisher JA. Hyperoxia‐induced hypocapnia: An underappreciated risk. Chest 128: 430‐433, 2005.
 147. Iversen PO, Standa M, Nicolaysen G. Marked regional heterogeneity in blood flow within a single skeletal muscle at rest and during exercise hyperaemia in the rabbit. Acta Physiol Scand 136: 17‐28, 1989.
 148. Jacoby I. Clostridial myositis, necrotizing fasciitis, and zygomycotic infections. In: Neuman TS, Thom SR, editors. The Physiology and Medicine of Hyperbaric Oxygen Therapy. Philadelphia: Saunders, 2008, p. 397‐418.
 149. Jarrett AS. Alveolar carbon dioxide tension at increased ambient pressures. J Appl Physiol 21: 158‐162, 1966.
 150. Johnston AJ, Steiner LA, Gupta AK, Menon DK. Cerebral oxygen vasoreactivity and cerebral tissue oxygen reactivity. Br J Anaesth 90: 774‐786, 2003.
 151. Keller H, Bühlmann AA. Deep diving and short decompression by breathing mixed gases. J Appl Physiol 20: 1267‐1270, 1965.
 152. Kerem D, Daskalovic YI, Arieli R, Shupak A. CO2 retention during hyperbaric exercise while breathing 40/60 nitrox. Undersea Hyperbaric Med 22: 339‐346, 1995.
 153. Kerem D, Melamed Y, Moran A. Alveolar Pco2 during rest and exercise in divers and non‐divers breathing O2 at 1 atm abs. Undersea Biomed Res 7: 17‐26, 1980.
 154. Kety SS. The theory and applications of the exchange of inert gas at the lungs and tissues. Pharmacol Rev 3: 1‐40, 1951.
 155. Kiesow LA. Hyperbaric inert gases and the hemoglobin‐oxygen equilibrium in red blood cells. Undersea Biomed Res 1: 29‐43, 1974.
 156. Kindwall EP. Compressed air work. In: Brubakk AO, Neuman TS, editors. Bennett and Elliott's Physiology and Medicine of Diving (5th ed). Edinburgh, UK: Saunders, 2003, p. 17‐28.
 157. Kindwall EP, Baz A, Lightfoot EN, Lanphier EH, Seireg A. Nitrogen elimination in man during decompression. Undersea Biomed Res 2: 285‐297, 1975.
 158. Kjellmer I, Lindberg I, Prerovsky I, Tønnesen H. The relation between blood flow in an isolated muscle measured with the Xe133 clearance and a direct recording technique. Acta Physiol Scand 69: 69‐78, 1967.
 159. Klein B, Kuschinsky W, Schrock H, Vetterlein F. Interdependency of local capillary density, blood flow, and metabolism in rat brains. Am J Physiol 251: H1333‐H1340, 1986.
 160. Koulouris NG, Valta P, Lavoie A, Corbeil C, Chassé M, Braidy J, Milic‐Emili J. A simple method to detect expiratory flow limitation during spontaneous breathing. Eur Respir J 8: 306‐313, 1995.
 161. Krichevsky IR, Kasarnovsky JS. Thermodynamical calculations of solubilities of nitrogen and hydrogen in water at high pressures. J Am Chem Soc 57: 2168‐2171, 1935.
 162. Kurss DI, Lundgren CEG, Påsche AJ. Effect of water temperature on vital capacity in head‐out immersion. In: Bachrach AJ, Matzen MM, editors. Underwater Physiology VII. Proceedings of the 5th Symposium on Underwater Physiology. Bethesda, MD: Federation of American Societies for Experimental Biology, 1981, p. 297‐301.
 163. Kvale PA, Davis J, Schroter RC. Effect of gas density and ventilatory pattern on steady‐state CO uptake by the lung. Respir Physiol 24: 385‐398, 1975.
 164. Kylstra JA, Paganelli CV, Lanphier EH. Pulmonary gas exchange in dogs ventilated with hyperbarically oxygenated liquid. J Appl Physiol 21: 177‐184, 1966.
 165. Lally DA, Zechman FW, Tracy RA. Ventilatory responses to exercise in divers and non‐divers. Resp Physiol 20: 117‐129, 1974.
 166. Lambertsen CJ, Ewing JH, Kough RH, Gould R, Stroud MW. Oxygen toxicity. Arterial and internal jugular blood gas composition in man during inhalation of air, 100% O2 and 2% CO2 in O2 at 3.5 atmospheres ambient pressure. J Appl Physiol 8: 255‐263, 1955.
 167. Lambertsen CJ, Idicula J. A new gas lesion syndrome in man, induced by “isobaric gas counterdiffusion”. J Appl Physiol 39: 434‐443, 1975.
 168. Lambertsen CJ, Kough RH, Cooper DY, Emmel GL, Loeschcke HH, Schmidt CF. Oxygen toxicity: Effects in man of oxygen inhalation at 1 and 3.5 atmospheres upon blood gas transport, cerebral circulation and cerebral metabolism. J Appl Physiol 5: 471‐86, 1953.
 169. Lambertsen CJ, Owen SG, Wendel H, Stroud MW, Lurie AA, Lochner W, Clark GF. Respiratory and circulatory control during exercise at 0.21 and 2.0 atmospheres inspired Po2. J Appl Physiol 14: 966‐982, 1959.
 170. Lango T, Morland T, Brubakk AO. Diffusion coefficients and solubility coefficients for gases in biological fluids: A review. Undersea Hyperb Med 23: 247‐272, 1996.
 171. Lanphier EH. Nitrogen‐Oxygen Mixture Physiology Phases 1 and 2 (Research Report 7‐55). Washington, DC: Navy Experimental Diving Unit, 1955.
 172. Lanphier EH. Nitrogen‐Oxygen Mixture Physiology Phases 4 and 6 (Research Report 7‐58). Washington, DC: Navy Experimental Diving Unit, 1958.
 173. Lanphier EH, Bookspan J. Carbon dioxide retention. In: Lundgren CEG, Miller JN, editors. The Lung at Depth. New York: Marcel Decker, 1999, p. 211‐236.
 174. Lassen NA. Blood flow of the cerebral cortex calculated from 85krypton‐beta‐clearance recorded over the exposed surface; evidence of inhomogeneity of flow. Acta Neurol Scand Suppl 14: 24‐28, 1965.
 175. Lee YC, Wu YC, Gerth WA, Vann RD. Absence of intravascular bubble nucleation in dead rats. Undersea Hyperb Med 20: 289‐296, 1993.
 176. Lin YC, Kakitsuba N, Watanabe DK, Mack GW. Influence of blood flow on cutaneous permeability to inert gas. J Appl Physiol 57: 1167‐1172, 1984.
 177. Linnarsson D, Hesser CM. Dissociated ventilatory and central respiratory responses to CO2 at raised N2 pressure. J Appl Physiol 45: 756‐761, 1978.
 178. Linnarsson D, Ostlund A, Lind F, Hesser CM. Hyperbaric bradycardia and hypoventilation in exercising med: Effects of ambient pressure and breathing gas. J Appl Physiol 87: 1428‐1432, 1999.
 179. Lippmann J, Mitchell SJ. Deeper into Diving (2nd ed). Melbourne, Victoria, Australia: Submariner Publications, 2005.
 180. Lumb AB. Nunn's Applied Respiratory Physiology (6th ed). Philadelphia: Elsevier, 2005.
 181. Lundgren C, Bergoe G, Olszowka A, Tyssebotn I. Tissue nitrogen elimination in oxygen‐breathing pigs is enhanced by fluorocarbon‐derived intravascular micro‐bubbles. Undersea Hyperb Med 32: 215‐226, 2005.
 182. Lundgren CEG. Respiratory function during simulated wet dives. Undersea Biomed Res 11: 139‐147, 1984.
 183. Lundgren CEG. Immersion effects. In: Lundgren CEG, Miller JN, editors. The Lung at Depth. New York: Marcel Decker, 1999, p. 91‐128.
 184. Lundgren CEG, Harabin A, Bennett PB, Van Liew HD, Thalmann ED. Gas physiology in diving. In: Fregly MJ, Blatteis CM, editors. Handbook of Physiology: Bethesda, MD: American Physiological Society, 1987, sect. 4, vol. II, chapt. 43, p. 999‐1022.
 185. Lundin G. Nitrogen elimination during oxygen breathing. Acta Physiol Scand 30 (Suppl 111): 130‐143, 1953.
 186. Macklem PT, Mead J. Resistance of central and peripheral airways measured by a retrograde catheter. J Appl Physiol 22: 395‐401, 1967.
 187. Mahon RT, Kerr S, Amundson D, Parrish JS. Immersion pulmonary edema in special forces combat swimmers. Chest 122: 383‐384, 2002.
 188. Maio DA, Farhi LE. Effect of gas density on mechanics of breathing. J Appl Physiol 23: 687‐693, 1967.
 189. Mapleson WW. Quantitative prediction of anesthetic concentrations. In: Papper EM, Kitz RJ, editors. Uptake and Distribution of Anesthetic Agents. New York: McGraw‐Hill, 1962, p. 104‐119.
 190. Marshall JM. Nitrogen narcosis in frogs and mice. Am J Physiol 166: 699‐710, 1951.
 191. McClaran SR, Wetter TJ, Pegelow DF, Dempsey JA. Role of expiratory flow limitation in determining lung volumes and ventilation during exercise. J Appl Physiol 86: 1357‐1366, 1999.
 192. McDonough PM, Hemmingsen EA. Bubble formation in crustaceans following decompression from hyperbaric gas exposure. J Appl Physiol 56: 513‐519, 1984.
 193. McMahon TJ, Moon RE, Luschinger BP, Carraway MS, Stone AE, Stolp BW, Gow AJ, Pawloski JR, Watke P, Singel DJ, Piantadosi CA, Stamler JS. Nitric oxide in the human respiratory cycle. Nature Med 8: 711‐717, 2002.
 194. Menn SJ, Sinclair RD, Welch BE. Effect of inspired CO2 up to 30mmHg on response of normal man to exercise. J Appl Physiol 28: 663‐671, 1970.
 195. Miller JN. Physiological limits to breathing dense gas. In: Lundgren CEG, Warkander DE, editors. Physiological and Human Engineering Aspects of Underwater Breathing Apparatus. Proceedings of the 40th Undersea and Hyperbaric Medical Society Workshop. Washington, DC: Undersea & Hyberbaric Medical Society, 1989, p. 5‐18.
 196. Mitchell SJ. Immersion pulmonary oedema. South Pacific Underwater Med Soc J 32: 200‐204, 2002.
 197. Mitchell SJ, Bennett MH. Clearance for diving and fitness for work. In: Neuman TS, Thom SR, editors. The Physiology and Medicine of Hyperbaric Oxygen Therapy. Philadelphia: Saunders, 2008, p. 65‐94.
 198. Mitchell SJ, Cronje F, Meintjes WAJ, Britz HC. Fatal respiratory failure during a technical rebreather dive at extreme pressure. Aviat Space Environ Med 78: 81‐86, 2007.
 199. Momson C. Report on the Use of Helium Oxygen Mixtures for Diving (Research Report 2). Washington, DC: Navy Experimental Diving Unit, 1939.
 200. Moon RE, Cherry AD, Stolp BW, Camporesi EM. Pulmonary gas exchange in diving. J Appl Physiol 106: 668‐677, 2009.
 201. Moon RE, Gorman DF. Treatment of decompression disorders. In: Brubakk AO, Neuman TS, editors. Bennett and Elliott's Physiology and Medicine of Diving (5th ed). Edinburgh, UK: Saunders, 2003, p. 600‐650.
 202. Moon RE, Gorman DF. Decompression sickness. In: Neuman TS, Thom SR, editors. The Physiology and Medicine of Hyperbaric Oxygen Therapy. Philadelphia: Saunders, 2008, p. 283‐320.
 203. Morrison JB, Florio JT, Butt WS. Observations after loss of consciousness underwater. Undersea Biomed Res 5: 179‐187, 1978.
 204. Moulder PV, Lancaster JR, Harrison RW, Michel SL, Snyder M, Thompson RG. Pulmonary arterial hyperoxia producing increased pulmonary vascular resistance. J Thorac Cardiovasc Surg 40: 588‐601, 1960.
 205. Mummery HJ, Stolp BW, Dear G deL, Doar PO, Natoli MJ, Boso AE, Archibold JD, Hobbs GW, El‐Moalem HE, Moon RE. Effects of age and exercise on physiological dead space during simulated dives at 2.8 atm abs. J Appl Physiol 94: 507‐517, 2003.
 206. Nairn JR, Power GG, Hyde RW, Forster RE, Lambertsen CJ, Dickson J. Diffusing capacity and pulmonary capillary blood flow at hyperbaric pressures. J Clin Invest 44: 1591‐1599, 1965.
 207. Naval Sea Systems Command. U.S. Navy Diving Manual. Arlington, VA: Naval Sea Systems Command, 2008.
 208. Navy Department. Diving Manual. Washington, DC: Navy Department, 1916.
 209. Neuman TS, Thom SR. The Physiology and Medicine of Hyperbaric Oxygen Therapy. Philadelphia: Saunders, 2008.
 210. Nishi RY, Lauckner GR. Development of the DCIEM 1983 Decompression Model for Compressed Air Diving (Report 84‐R‐88). Downsview, Ontario, Canada: Defence and Civil Institute of Environmental Medicine, 1984.
 211. Novotny JA, Mayers DL, Parsons Y‐FJ, Survanshi SS, Weathersby PK, Homer LD. Xenon kinetics in muscle are not explained by a model of parallel perfusion‐limited compartments. J Appl Physiol 68: 876‐890, 1990.
 212. Ohta Y, Farhi LE. Cerebral gas exchange: Perfusion and diffusion limitations. J Appl Physiol 46: 1164‐1168, 1979.
 213. Paiva M, Engel LA. Pulmonary interdependence of gas transport. J Appl Physiol 47: 296‐305, 1979.
 214. Peacher DF, Pecorella SRH, Freiberger JJ, Natoli MJ, Schinazi EA, Doar O, Boso AE, Walker AJ, Gill M, Kernagis D, Ugucionni D, Moon RE. Effects of hyperoxia on ventilation and pulmonary hemodynamics during immersed prone exercise at 4.7 atm abs: Possible implications for immersion pulmonary edema. J Appl Physiol 109: 68‐78, 2010.
 215. Pedley TJ, Schroter RC, Sudlow MF. The prediction of pressure drop and variation of resistance within the human bronchial airways. Resp Physiol 9: 387‐405, 1970.
 216. Pendergast DR, Lindholm P, Wylegala J, Warkander D, Lundgren CEG. Effects of respiratory muscle training on respiratory sensitivity in SCUBA divers. Undersea Hyper Med 33: 447‐453, 2006.
 217. Perl W, Rackow H, Salanitre E, Wolf GL, Epstein RM. Intertissue diffusion effect for inert fat‐soluble gases. J Appl Physiol 20: 621‐627, 1965.
 218. Piantadosi C. Pulmonary gas exchange, oxygen transport, and tissue oxygenation. In: Neuman TS, Thom SR, editors. The Physiology and Medicine of Hyperbaric Oxygen Therapy. Philadelphia: Saunders, 2008, p. 133‐158.
 219. Piiper J, Scheid P. Diffusion and convection in intrapulmonary gas mixing. In: Farhi LE, Tenney SM, editors. Handbook of Physiology. Bethesda, MD: American Physiological Society, 1987, sect. 3, vol. IV, chapt. 4, p. 51‐70.
 220. Poon CS. Ventilatory control in hypercapnia and exercise: Optimization hypothesis. J Appl Physiol 62: 2447‐2459, 1987.
 221. Poon CS. Effects of inspiratory elastic load on respiratory control in hypercapnia and exercise. J Appl Physiol 66: 2400‐2406, 1989.
 222. Poon CS. Effects of inspiratory resistive load on respiratory control in hypercapnia and exercise. J Appl Physiol 66: 2391‐2399, 1989.
 223. Pope H, Holloway R, Campbell EJM. The effects of elastic and resistive loading of inspiration on the breathing of conscious man. Respir Physiol 4: 363‐372, 1968.
 224. Prausnitz JM. Molecular Thermodynamics of Fluid‐Phase Equilibria. Englewood Cliffs, NJ: Prentice‐Hall, 1969.
 225. Prefaut C, Dubois F, Roussos C, Amaral‐Marques R, Macklem PT, Ruff F. Influence of immersion to the neck in water on airway closure and distribution of perfusion in man. Respir Physiol 37: 313‐323, 1979.
 226. Prefaut C, Lupi‐H E, Anthonisen NR. Human lung mechanics during water immersion. J Appl Physiol 40: 320‐323, 1976.
 227. Puy RJM, Hyde RW, Fisher AB, Clark JM, Dickson J, Lambertsen CJ. Alterations in the pulmonary capillary bed during early O2 toxicity in man. J Appl Physiol 24: 537‐543, 1968.
 228. Reeves RB, Morin RA. Pressure increases oxygen affinity of whole blood and erythrocyte suspensions. J Appl Physiol 61: 486‐494, 1986.
 229. Ringqvist T. The ventilatory capacity in healthy subjects. Scand J Clin Lab Invest 18 (Suppl 88): 1‐179, 1966.
 230. Saltzman HA. Rational normobaric and hyperbaric oxygen therapy. Ann Intern Med 67: 843‐852, 1967.
 231. Saltzman HA, Salzano JV, Blenkarn GD, Kylstra JA. Effects of pressure on ventilation and gas exchange in man. J Appl Physiol 30: 443‐449, 1971.
 232. Salzano J, Rausch DC, Saltzman HA. Cardiorespiratory responses to exercise at a simulated seawater depth of 1,000 feet. J Appl Physiol 28: 34‐41, 1970.
 233. Salzano JV, Camporesi EM, Stolp BW, Moon RE. Physiological responses to exercise at 47 and 66 atm abs. J Appl Physiol 57: 1055‐1068, 1984.
 234. 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.
 235. Sejrsen P. Shunting by diffusion of gas in skeletal muscle and brain. In: Johansen K, Burggren W, editors. Cardiovascular Shunts: Phylogenic, Ontogenic, and Clinical Aspects. Proceedings of the Alfred Benzon Symposium 21. Copenhagen, Denmark: Munksgaard, 1985, p. 452‐462.
 236. Sejrsen P, Tønnesen KH. Inert gas diffusion method for measurement of blood flow using saturation techniques. Circ Res 22: 679‐693, 1968.
 237. Sejrsen P, Tønnesen KH. Shunting by diffusion of inert gas in skeletal muscle. Acta Physiol Scand 86: 82‐91, 1972.
 238. Shepard RH, Campbell EJM, Martin HB, Enns T. Factors affecting the pulmonary dead space as determined by single breath analysis. J Appl Physiol 11: 241‐244, 1957.
 239. Shephard RJ. The maximum sustained voluntary ventilation in exercise. Clin Sci 32: 167‐76, 1967.
 240. Sherman D, Eilender E, Shefer A, Kerem D. Ventilatory and occlusion‐pressure responses to hypercapnia in divers and non‐divers. Undersea Biomed Res 7: 61‐74, 1980.
 241. Shupak A, Weiler‐Ravell D, Adir Y, Daskalovic YI, Ramon Y, Kerem D. Pulmonary oedema induced by strenuous swimming; a field study. Respir Physiol 121: 25‐31, 2000.
 242. Slade JB Jr, Hattori T, Ray CS, Bove AA, Cianci P. Pulmonary edema associated with scuba diving: Case reports and review. Chest 120: 1686‐1694, 2001.
 243. Slutsky AS. Gas mixing by cardiogenic oscillations: A theoretical quantitative analysis. J Appl Physiol 51: 1287‐1293, 1981.
 244. Smart DR, Bennett MH, Mitchell SJ. Transcutaneous oximetry, problem wounds and hyperbaric oxygen therapy. Diving Hyperb Med 36: 72‐86, 2006.
 245. Srinivasan RS, Gerth WA, Powell MR. Mathematical models of diffusion‐limited gas bubble dynamics in tissue. J Appl Physiol 86: 732‐741, 1999.
 246. Srinivasan RS, Gerth WA, Powell MR. Mathematical model of diffusion‐limited gas bubble dynamics in unstirred tissue with finite volume. Ann Biomed Eng 30: 232‐246, 2002.
 247. Srinivasan RS, Gerth WA, Powell MR. Mathematical model of diffusion‐limited evolution of multiple gas bubbles in tissue. Ann Biomed Eng 31: 471‐481, 2003.
 248. Stolp BW, Moon RE, Salzano JV, Camporesi EM. 2,3 DPG levels during saturation diving to 650 msw. Undersea Biomed Res 12: 21, 1985.
 249. Survanshi SS, Parker EC, Gummin DD, Flynn ET, Toner CB, Temple DJ, Ball R, Homer LD. Human Decompression Trial with 1.3 ATA Oxygen in Helium (Technical Report 98‐09). Bethesda, MD: Naval Medical Research Institute, 1998.
 250. Taunton JE, Banister EW, Patrick TR, Oforsagd P, Duncan WR. Physical work capacity in hyperbaric environments and conditions of hyperoxia. J Appl Physiol 28: 421‐427, 1970.
 251. Taylor NAS, Morrison JB. Effects of breathing‐gas pressure on pulmonary function and work capacity during immersion. Undersea Biomed Res 17: 413‐428, 1990.
 252. Taylor NAS, Morrison JB. Static respiratory muscle work during immersion with positive and negative respiratory loading. J Appl Physiol 87: 1397‐1403, 1999.
 253. Teorell T. Kinetics of distribution of substances administered to the body, part II: The intravascular modes of administration. Arch Int Pharmacodyn Ther 57: 226‐240, 1937.
 254. Tepper RS, Lightfoot EN, Baz A, Lanphier EH. Inert gas transport in the microcirculation: Risk of isobaric supersaturation. J Appl Physiol 46: 1157‐1163, 1979.
 255. Thalmann ED. Phase II Testing of Decompression Algorithms for Use in the U.S. Navy Underwater Decompression Computer (Technical Report 1‐84). Panama City, FL: Navy Experimental Diving Unit, 1984.
 256. Thalmann ED. Development of a Decompression Algorithm for Constant 0.7 ATA Oxygen Partial Pressure in Helium Diving (Technical Report 1‐85). Panama City, FL: Navy Experimental Diving Unit, 1985.
 257. Thalmann ED. Rebreather basics. In: Richardson D, editor. Proceedings of Rebreather Forum 2.0. Redondo Beach, CA: Diving Science and Technology, 1996, p. 23‐30.
 258. Thalmann ED, Parker EC, Survanshi SS, Weathersby PK. Improved probabilistic decompression model risk predictions using linear‐exponential kinetics. Undersea Hyperb Med 24: 255‐274, 1997.
 259. Thalman ED, Sponholtz DK, Lundgren GEG. Effects of immersion and static lung loading on submerged exercise at depth. Undersea Biomed Res 6: 259‐290, 1979.
 260. Thalmann ED, Survanshi SS, Flynn ET. Direct comparison of the effects of He, N2, and wet or dry conditions on the 60 fsw no‐decompression limit [abstract]. Undersea Biomed Res 16: 67, 1989.
 261. Thavasothy M, Broadhead M, Elwell C, Peters M, Smith M. A comparison of cerebral oxygenation as measured by the NIRO 300 and the INVOS 5100 near‐infrared spectrophotometers. Anesthesia 57: 999‐1006, 2002.
 262. Tikuisis P, Gault KA, Nishi RY. Prediction of decompression illness using bubble models. Undersea Hyperb Med 21: 129‐143, 1994.
 263. Tikuisis P, Gerth WA. Decompression theory. In: Brubakk AO, Neuman TS, editors. Bennett and Elliott's Physiology and Medicine and Diving (5th ed). Edinburgh: Saunders, 2003, p. 419‐454.
 264. Tikuisis P, Nishi RY, Weathersby PK. Use of the maximum likelihood method in the analysis of chamber air dives. Undersea Biomed Res 15: 301‐313, 1988.
 265. Trytko B, Mitchell SJ. Extreme survival: A deep technical diving accident. South Pacific Underwater Med Soc J 35: 23‐7, 2005.
 266. Van Der Aue OE, Kellar RJ, Brinton ES, Barron G, Gilliam HD, Jones RJ. Calculation and Testing of Decompression Tables for Air Dives Employing the Procedure of Surface Decompression and the Use of Oxygen (Research Report 13‐51). Washington, DC: Navy Experimental Diving Unit, 1951.
 267. Van Liew HD. Mechanical and physical factors in lung function during work in dense environments. Undersea Biomed Res 10: 255‐264, 1983.
 268. Van Liew HD. Simulation of the dynamics of decompression sickness bubbles and the generation of new bubbles. Undersea Biomed Res 18: 333‐345, 1991.
 269. Van Liew HD, Bishop B, Walder‐D P, Rahn H. Effects of compression on composition and absorption of tissue gas pockets. J Appl Physiol 20: 927‐933, 1965.
 270. Van Liew HD, Conkin J, Burkard ME. The oxygen window and decompression bubbles: Estimates and significance. Aviat Space Environ Med 64: 859‐865, 1993.
 271. Van Liew HD, Hlastala MP. Influence of bubble size and blood perfusion on absorbtion of gas bubbles in tissue. Resp Physiol 7: 111‐121, 1969.
 272. Van Liew HD, Thalmann ED, Sponholtz DK. Hindrance to diffusive gas mixing in the lung in hyperbaric environments. J Appl Physiol 51: 243‐247, 1981.
 273. Verstappen FT, Bernards JA, Kreuzer F. Effects of pulmonary gas embolism on circulation and respiration in the dog, part II: Effects on respiration. Pflugers Arch 368: 97‐104, 1977.
 274. von Kummer R, Herold S. Hydrogen clearance method for determining local cerebral blood flow, part I: Spatial resolution. J Cereb Blood Flow Metab 6: 486‐491, 1986.
 275. von Kummer R, von Kries F, Herold S. Hydrogen clearance method for determining local cerebral blood flow, part II: Effect of heterogeneity in cerebral blood flow. J Cereb Blood Flow Metab 6: 492‐498, 1986.
 276. Vorosmarti J Jr, Barnard EE, Williams J, de GH. Nitrogen elimination during steady‐state hyperbaric exposures. Undersea Biomed Res 5: 243‐252, 1978.
 277. Wagner PD. Insensitivity of Vo2max to hemoglobin‐P50 at sea level and altitude. Respir Physiol 107: 205‐212, 1997.
 278. Wagner PD, Laravuso RB, Uhl RR, West JB. Continuous distributions of ventilation‐perfusion ratios in normal subjects breathing air and 100% O2. J Clin Invest 54: 54‐68, 1974.
 279. Wagner PD, Saltzman HA, West JB. Measurement of continuous distributions of ventilation‐perfusion ratios: Theory. J Appl Physiol 36: 588‐589, 1974.
 280. Ward CA, Balakrishnan A, Hooper FC. On the thermodynamics of nucleation in weak gas‐liquid solutions. J Basic Eng 92: 695‐704, 1970.
 281. Warkander DE. Development of a scrubber gauge for closed‐circuit diving [abstract]. Undersea Hyper Med 34: 251, 2007.
 282. Warkander DE, Lundgren CEG. Effects of graded combinations of resistance and elastance on divers’ respiratory performance. Undersea Hyper Med 21, (Suppl): 96‐97, 1994.
 283. Warkander DE, Lundgren CEG. Effects of combinations of resistance and elastance on divers’ respiratory performance during exposure to a negative static lung load [abstract]. Undersea Hyper Med 22 (Suppl): 71, 1995.
 284. Warkander DE, Nagasawa GK, Lundgren CEG. Effects of inspiratory and expiratory resistance in divers’ breathing apparatus. Undersea Hyper Med 28: 63‐73, 2001.
 285. Warkander DE, Norfleet WT, Nagasawa GK, Lundgren CE. CO2 retention with minimal symptoms but severe dysfunction during wet simulated dives to 6.8 atm abs. Undersea Biomed Res 17: 515‐523, 1990.
 286. Warkander DE, Norfleet WT, Nagasawa GK, Lundgren CEG. Physiologically and subjectively acceptable breathing resistance in divers’ breathing gear. Undersea Biomed Res 19: 427‐445, 1992.
 287. Wasserman K. Testing regulation of ventilation with exercise. Chest 70 (Suppl): 173‐178, 1975.
 288. Weathersby PK, Mendenhall KG, Barnard EEP, Homer LD, Survanshi S, Vieras F. Distribution of xenon gas exchange rates in dogs. J Appl Physiol 50: 1325‐1336, 1981.
 289. Weathersby PK, Meyer P, Flynn ET, Homer LD, Survanshi S. Nitrogen gas exchange in the human knee. J Appl Physiol 61: 1534‐1545, 1986.
 290. Weibe R, Gaddy VL, Heins C. The compressibility isotherms of helium at temperatures from −70 to 200°C and at pressures to 1000 atmospheres. J Am Chem Soc 53: 1721‐1725, 1931.
 291. Weibel ER, Sapoval B, Filoche M. Design of peripheral airways for efficient gas exchange. Respir Physiol Neurobiol 148: 3‐21, 2005.
 292. Whalen RE, Saltzman HA, Holloway DH, McIntosh HD, Sieker HO, Brown IW. Cardiovascular and blood gas responses to hyperbaric oxygenation. Am J Cardiol 15: 638‐646, 1965.
 293. Whitelaw WAJ, Derenne JP, Milic‐Emili J. Occlusion pressure as a measure of respiratory centre output in conscious man. Respir Physiol 23: 181‐199, 1975.
 294. Willmon TL, Behnke AR. Nitrogen elimination and oxygen absorbtion at high barometric pressure. Am J Physiol 131: 633‐638, 1940.
 295. Wilmshurst PT, Crowther A, Nuri M, Webb‐Peploe MM. Cold‐induced pulmonary oedema in scuba divers and swimmers and subsequent development of hypertension. Lancet i: 62‐65, 1989.
 296. Wilmshurst PT, Nuri M, Crowther A, Webb‐Peploe MM. Forearm vascular responses in subjects who develop recurrent pulmonary edema when scuba diving; a new syndrome. Br Heart J 45 (Suppl): A349, 1982.
 297. Wolpers HG, Hoeft A, Korb H, Lichtlen PR, Hellige G. Transport of inert gases in mammalian myocardium: Comparison with a convection‐diffusion model. Am J Physiol 259: H167‐H173, 1990.
 298. Wood LDH, Bryan AC. Effect of increased ambient pressure on flow‐volume curve of the lung. J Appl Physiol 27: 4‐8, 1969.
 299. Wood LDH, Bryan AC. Exercise ventilatory mechanics at increased ambient pressure. J Appl Physiol 44: 231‐237, 1978.
 300. Wood LDH, Bryan AC, Bau SK, Weng TR, Levison H. Effect of increased gas density on pulmonary gas exchange in man. J Appl Physiol 41: 206‐210, 1976.
 301. Workman RD. Calculation of Decompression Schedules for Nitrogen‐Oxygen and Helium‐Oxygen Dives (Research Report 6‐65). Washington, DC: Navy Experimental Diving Unit, 1965.
 302. Young C, D'Aoust BG. Factors determining temporal pattern of isobaric supersaturation. J Appl Physiol 51: 852‐857, 1981.
 303. Yount DE, Hoffman DC. On the use of a bubble formation model to calculate diving tables. Aviat Space Environ Med 57: 149‐156, 1986.
 304. Yount DE, Lally DA. On the use of oxygen to facilitate decompression. Aviat Space Environ Med 51: 544‐550, 1980.
 305. Yount DE, Maiken EB, Baker EC. Implications of the varying permeability model for reverse dive profiles. In: Lang MA, Lehner CE, editors. Proceedings of the Reverse Dive Profiles Workshop. Washington, DC: Smithsonian Institution, 2000, p. 29‐60.
 306. Zechman F, Hall FG, Hull WE. Effects of graded resistance to tracheal air flow in man. J Appl Physiol 10: 356‐362, 1957.

Contact Editor

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

* Required Field

How to Cite

David J. Doolette, Simon J. Mitchell. Hyperbaric Conditions. Compr Physiol 2010, 1: 163-201. doi: 10.1002/cphy.c091004