Partial pressure
Partial pressure is the hypothetical pressure that a single gas component in a mixture would exert if it occupied the entire volume of the container alone, at the same temperature and volume as the mixture.[1] This concept forms the basis of Dalton's law of partial pressures, which states that for a mixture of non-reacting ideal gases, the total pressure is equal to the sum of the partial pressures of each individual gas.[2] Formulated by English chemist John Dalton in 1802, the law assumes that gases do not interact chemically and behave independently under constant temperature and volume conditions.[3] The partial pressure of a gas, denoted as P_i, is quantitatively determined by multiplying the total pressure P of the mixture by the mole fraction x_i of that gas, where x_i = \frac{n_i}{n_{\text{[total](/page/Total)}}} and n_i is the number of moles of the gas, yielding P_i = x_i \cdot P.[4] This relationship holds for ideal gases and is fundamental in gas stoichiometry, enabling calculations of gas compositions in reactions and mixtures.[5] In practical terms, partial pressures are crucial for understanding gas behavior in diverse fields, including atmospheric science, where they describe the contributions of nitrogen (about 78%) and oxygen (about 21%) to sea-level air pressure of 760 mmHg, resulting in partial pressures of approximately 593 mmHg and 160 mmHg, respectively.[6] Beyond chemistry, partial pressures play a vital role in physiology, particularly in respiratory gas exchange, where the partial pressure of oxygen (P_{O_2}) in alveolar air drives diffusion into the bloodstream, and elevated P_{CO_2} in tissues promotes carbon dioxide release.[7] In clinical settings, monitoring arterial partial pressures via blood gas analysis assesses oxygenation, ventilation, and acid-base balance, guiding treatments for conditions like respiratory failure.[8] Additionally, the concept is essential in applications such as scuba diving, where increased total pressure at depth raises partial pressures of inert gases, risking decompression sickness, and in industrial processes like gas separation and combustion efficiency optimization.[9]Basic Principles
Definition
Partial pressure is defined as the hypothetical pressure exerted by an individual gas component in a mixture if that gas alone occupied the entire volume of the container at the same temperature as the mixture.[10] This measure assumes that the gases in the mixture do not chemically interact and behave independently, allowing each to contribute proportionally to the total pressure based on its mole fraction.[10] The concept of partial pressure was first formulated by English chemist and physicist John Dalton in 1801, as part of his investigations into gas behavior that laid foundational groundwork for his atomic theory of matter.[11] Dalton's work emphasized that the pressure contributions from each gas arise from the repulsive forces between their constituent particles, viewing mixtures as mechanical rather than chemical combinations.[11] This 1801 insight, later formalized as Dalton's law of partial pressures, provided a key empirical basis for understanding gaseous systems.[11] To intuitively grasp partial pressure, consider a mixture of gases as akin to multiple individuals in a shared space, each generating sound independently that collectively determines the total noise level, without one altering the output of another. In a practical example, Earth's atmosphere at sea level has a total pressure of about 101 kPa, with oxygen comprising roughly 21% of dry air, yielding an oxygen partial pressure of approximately 21 kPa.[12]Notation and Units
In scientific literature, the partial pressure of a gas component i in a mixture is commonly denoted by p_i or P_i, where the subscript i specifies the identity of the component, such as the chemical symbol or formula of the gas. The total pressure of the mixture is typically represented by p or P, with a common convention using uppercase P for the total pressure and lowercase p_i for partial pressures to highlight the distinction. The International Union of Pure and Applied Chemistry (IUPAC) recommends the lowercase notation, defining the partial pressure as p_i = x_i p, where x_i is the mole fraction of component i and p is the total pressure.[13][14] Partial pressures are quantified using the same units as pressure measurements. The SI unit is the pascal (Pa), equivalent to one newton per square meter (N/m²), which provides a standardized metric for precise calculations in thermodynamics and gas dynamics. In practice, especially in chemistry and physiology, other units are prevalent, including the atmosphere (atm), bar, millimeter of mercury (mmHg), and torr; for instance, normal atmospheric pressure is 1 atm, or exactly 101325 Pa. Conversions between these units are essential: 1 bar = 100000 Pa, 1 mmHg = 133.322 Pa, and 1 torr ≈ 133.322 Pa (with 1 torr defined as exactly 1/760 atm)./13:_States_of_Matter/13.04:_Pressure_Units_and_Conversions) In equations involving gas mixtures, subscripts clearly denote specific species, such as p_{\ce{[O2](/page/O2)}} for the partial pressure of oxygen or p_{\ce{[N2](/page/Nitrogen)}} for nitrogen, ensuring unambiguous representation in formulas like those for equilibrium constants or solubility. The torr unit originated from the experiments of Italian physicist Evangelista Torricelli in 1643, who developed the mercury barometer to measure atmospheric pressure as the height of a mercury column, with one torr corresponding to the pressure exerted by 1 mm of mercury at 0°C. This historical unit remains widely used in vacuum technology and respiratory physiology due to its direct link to barometric measurements.[15][16]Dalton's Law
Dalton's law of partial pressures states that in a mixture of non-reacting gases, the total pressure exerted by the mixture is equal to the sum of the partial pressures that each individual gas would exert if it occupied the same volume alone at the same temperature.[17] The partial pressure of a component gas P_i is given by P_i = x_i P_{\text{total}}, where x_i is the mole fraction of that gas in the mixture, defined as the ratio of the number of moles of the component to the total number of moles of all gases.[18] This law can be derived from the kinetic molecular theory of gases, which posits that gas pressure arises from the random collisions of molecules with the container walls. In a mixture, molecules of each gas species collide independently with the walls, and the total pressure is the sum of contributions from each species, proportional to their respective number densities (moles per unit volume) since the average kinetic energy per molecule is the same for all gases at a given temperature./Physical_Properties_of_Matter/States_of_Matter/Properties_of_Gases/Gas_Laws/Daltons_Law_(Law_of_Partial_Pressures)) The law assumes that the gases behave ideally, meaning they are composed of point masses with no intermolecular forces, leading to independent contributions to pressure; it holds well under conditions of low pressure and high temperature where these assumptions are valid.[18] John Dalton formulated the law in 1801 based on his meteorological observations, particularly experiments measuring the pressure of water vapor in air, where he found that the vapor's pressure remained unchanged regardless of the presence of dry air, confirming the additive nature of pressures in mixtures.[17] In real gases, deviations from Dalton's law occur at high pressures due to intermolecular attractions and the finite volume of molecules, which can be partially accounted for using equations like the van der Waals equation for mixtures, though the law remains a good approximation for most practical scenarios involving dilute gases.[19]Gas Mixture Properties
Ideal Gas Approximations
In ideal gas mixtures, the composition is quantified using the mole fraction, defined as the ratio of the number of moles of a specific component n_i to the total number of moles in the mixture n_{\text{total}}, expressed as x_i = \frac{n_i}{n_{\text{total}}}.[20] This mole fraction directly relates to partial pressure through the relation P_i = x_i P_{\text{total}}, where P_i is the partial pressure of component i and P_{\text{total}} is the total pressure of the mixture.[21] This connection stems from Dalton's law, which posits that the total pressure is the sum of individual partial pressures in non-interacting ideal gases.[22] For each component in an ideal gas mixture, the partial pressure follows the ideal gas law independently: P_i V = n_i R T, where V is the shared volume, R is the universal gas constant, and T is the temperature.[22] Substituting the mole fraction yields P_i = x_i \frac{n_{\text{total}} R T}{V} = x_i P_{\text{total}}, confirming the proportionality. Summing over all components gives P_{\text{total}} = \sum P_i = \left( \sum n_i \right) \frac{R T}{V} = n_{\text{total}} \frac{R T}{V}, illustrating how the total pressure arises additively without intermolecular interactions affecting the behavior.[23] This framework applies to both binary and multicomponent mixtures. For instance, in dry air approximated as 78% nitrogen (N₂), 21% oxygen (O₂), and 1% argon (Ar) by mole fraction, at a total pressure of 1 atm (760 torr), the partial pressure of N₂ is $0.78 \times 760 = 593 torr, O₂ is $0.21 \times 760 = 160 torr, and Ar is $0.01 \times 760 = 8 torr.[24] These values highlight how partial pressures reflect the proportional contributions of each gas to the overall mixture pressure under ideal conditions. Thermodynamically, ideal gas mixtures exhibit additive properties for extensive state functions, with no enthalpy or internal energy of mixing. The total internal energy U is the sum U = \sum n_i u_i(T), where u_i(T) is the molar internal energy of pure component i depending only on temperature, and similarly for enthalpy H = \sum n_i h_i(T). This additivity simplifies calculations for processes like heating or compression, as changes in these properties depend solely on the individual components' temperature dependencies without cross-interactions.[25]Partial Volume and Amagat's Law
Amagat's law, also known as the law of additive volumes, states that the total volume of an ideal gas mixture at constant temperature and pressure equals the sum of the partial volumes of its individual components. The partial volume V_i of each component i is defined as the volume that component would occupy alone under the same total pressure P and temperature T as the mixture. Mathematically, this is expressed as V = \sum_i V_i, where V_i = \frac{n_i R T}{P}, with n_i denoting the moles of component i and R the universal gas constant.[26] This formulation connects directly to partial pressure via the ideal gas law. The partial pressure p_i of component i satisfies p_i V = n_i R T, so rearranging yields V_i = \frac{p_i V}{P}, indicating that each partial volume is the total volume scaled by the mole fraction (since p_i / P = x_i).[26] The law originated from experiments by French physicist Émile-Hilaire Amagat on the compressibility of gases under high pressures. In 1880, Amagat published findings from measurements up to several thousand atmospheres, leading to the law of additive volumes for ideal gas mixtures.[27] Amagat's law complements Dalton's law by offering a volume-additive viewpoint at fixed pressure and temperature, in contrast to Dalton's pressure-additive approach at fixed volume. Both describe ideal gas mixtures but apply to distinct conditions, enabling consistent predictions of mixture properties from either perspective.[26] A representative example is the mixing of 1 mole of hydrogen and 1 mole of oxygen at 298 K and 1 atm total pressure. Each gas alone would occupy 24.5 L (from the ideal gas law), so the mixture volume is 49.0 L with no deviation upon ideal mixing.[28]Related Thermodynamic Concepts
Vapor Pressure
Vapor pressure refers to the partial pressure exerted by a vapor in thermodynamic equilibrium with its liquid or solid phase in a closed system at a given temperature./Physical_Properties_of_Matter/States_of_Matter/Properties_of_Liquids/Vapor_Pressure) This equilibrium arises when the rate of evaporation equals the rate of condensation, resulting in a constant pressure attributable solely to the vapor component.[29] In mixtures containing non-condensable gases, the vapor pressure represents the partial pressure of the condensable component, distinct from the contributions of other gases.[30] The vapor pressure of a substance increases with temperature due to enhanced molecular kinetic energy, which favors evaporation over condensation. This temperature dependence is quantitatively described by the Clausius-Clapeyron equation: \frac{d \ln P_\text{vap}}{dT} = \frac{\Delta H_\text{vap}}{R T^2} where P_\text{vap} is the vapor pressure, T is the absolute temperature, \Delta H_\text{vap} is the enthalpy of vaporization, and R is the gas constant./Physical_Properties_of_Matter/States_of_Matter/Phase_Transitions/Clausius-Clapeyron_Equation) The equation highlights the exponential relationship between vapor pressure and temperature, explaining why boiling occurs when vapor pressure equals the surrounding total pressure. A representative example is water, whose saturation vapor pressure at 25°C is 3.17 kPa (23.8 mmHg).[31] In atmospheric contexts, this vapor pressure plays a key role in humidity metrics; relative humidity (RH) is defined as the ratio of the actual partial pressure of water vapor (P_\text{vap}) to the saturation vapor pressure (P_\text{sat}) at that temperature, expressed as RH = (P_\text{vap} / P_\text{sat}) \times 100\%.[32] For instance, at 25°C, an RH of 50% corresponds to a water vapor partial pressure of 1.585 kPa. In moist air, the total atmospheric pressure (P_\text{total}) is the sum of the partial pressure of dry air (P_\text{dry}) and the partial pressure of water vapor (P_\text{vap}), following Dalton's law of partial pressures: P_\text{total} = P_\text{dry} + P_\text{vap}.[33] This distinction is crucial, as P_\text{vap} can approach P_\text{sat} near saturation, potentially leading to condensation if exceeded. Vapor pressure is measured using static or dynamic techniques. Static methods involve equilibrating the sample in a closed vessel and directly gauging the pressure, often with manometers to avoid contamination. Dynamic methods, such as gas carrier techniques, pass a dry inert gas over the liquid and measure the increase in humidity or pressure downstream. A specific static variant is the isoteniscope method, which uses a capillary tube to trap a vapor bubble and indirectly determines pressure by balancing it against a reference manometer, minimizing sample loss and composition changes.Henry's Law
Henry's law describes the solubility of a gas in a liquid, stating that at a constant temperature, the concentration of the dissolved gas is directly proportional to its partial pressure in the gas phase above the liquid. This relationship arises from the equilibrium between the gas and the dissolved species, where the amount of gas absorbed increases linearly with the applied pressure. The law is expressed in its basic solubility form as C = k_H P_i where C is the molar concentration of the dissolved gas, P_i is the partial pressure of the gas, and k_H is the Henry's law constant, which is specific to the gas-liquid pair and temperature.[34][35] The law originates from observations by William Henry in 1803, who found that the quantity of gas dissolved in water under pressure is proportional to the degree of compression, marking an early quantitative study of gas absorption. Henry's law appears in several equivalent forms depending on the choice of variables, reflecting different conventions for the constant. The pressure-based form uses concentration as C = k P, while the mole fraction-based form is often x_i = K P_i (or inversely P_i = H x_i), where x_i is the mole fraction of the solute in the liquid; solubility-based variants express the ratio of gas volume dissolved to gas volume at standard pressure. These formulations are unified under IUPAC recommendations, which define eight variants to ensure consistency in thermodynamic applications, with the solubility constant H^s_{cp} = C / P_i being commonly used for aqueous systems.[34][36] The Henry's constant k_H exhibits a strong temperature dependence, generally decreasing as temperature rises, which reduces gas solubility—for most gases, higher temperatures lead to lower dissolution rates due to the exothermic nature of the solvation process. This effect is critical in practical scenarios, such as carbonated beverages where elevated CO₂ partial pressure at bottling achieves high dissolution, but warming or pressure release causes effervescence as solubility drops. Similarly, in oxygenation processes like water aeration for environmental remediation, maintaining appropriate partial pressures compensates for temperature-induced solubility limits to enhance oxygen transfer. Pressure effects beyond the linear regime are minimal at low partial pressures, but the law assumes ideal gas behavior.[1][37] As a limiting law analogous to the ideal gas law, Henry's law holds primarily for dilute solutions where solute concentrations are low and interactions are negligible, and for low partial pressures where gas ideality applies; significant deviations occur at high concentrations due to non-ideal solute-solvent interactions or at elevated pressures from compressibility effects. These limitations restrict its use to scenarios like sparingly soluble gases in solvents, ensuring accuracy in predictions of equilibrium partitioning.[34][38]Equilibrium in Gas Reactions
In gas-phase reactions at equilibrium, the equilibrium constant K_p is defined in terms of the partial pressures of the reactants and products, expressed as K_p = \prod (P_i)^{\nu_i}, where P_i is the partial pressure of species i and \nu_i are the stoichiometric coefficients (positive for products and negative for reactants).[39] This formulation arises because partial pressures directly reflect the effective concentration of gases in mixtures under ideal conditions, influencing the position of equilibrium.[40] The relationship between K_p and the concentration-based equilibrium constant K_c for gas reactions is given by K_p = K_c (RT)^{\Delta n}, where R is the gas constant, T is the absolute temperature, and \Delta n is the change in the number of moles of gas (\Delta n = \sum \nu_i for products minus reactants).[41] This connection highlights how partial pressures scale with total pressure and temperature, affecting equilibrium shifts when \Delta n \neq 0.[42] Thermodynamically, the influence of partial pressures on equilibrium stems from the Gibbs free energy change, \Delta G = \Delta G^\circ + RT \ln Q_p, where Q_p is the reaction quotient analogous to K_p but using instantaneous partial pressures, and equilibrium occurs when \Delta G = 0 and Q_p = K_p.[43] At equilibrium, \Delta G^\circ = -RT \ln K_p, linking standard free energy to partial pressure-based constants.[44] A key example is the Haber-Bosch process for ammonia synthesis, N_2 + 3H_2 \rightleftharpoons 2NH_3, where \Delta n = -2, so increasing total pressure raises partial pressures of reactants and shifts equilibrium toward products per Le Chatelier's principle, enhancing yield despite the exothermic nature favoring lower temperatures.[45] Industrial conditions typically use 200–300 atm and 400–500°C with iron catalysts to balance kinetics and thermodynamics.[46] In industrial catalysis, partial pressures guide reactor design for processes like ammonia or methanol synthesis, where high reactant partial pressures maximize K_p-driven conversions.[47] For combustion equilibria, such as in gas turbines, partial pressures of oxygen and fuels determine species distributions at high temperatures, influencing efficiency and emissions via K_p for reactions like CO + \frac{1}{2}O_2 \rightleftharpoons CO_2).Applications
Underwater Diving
In underwater diving, the partial pressures of gases in breathing mixtures increase with depth due to the ambient pressure, calculated as 1 atm at the surface plus 0.1 atm per meter of seawater depth.[48] Divers commonly use air, consisting of approximately 21% oxygen and 79% nitrogen, where the partial pressure of each gas is its fractional concentration multiplied by the total absolute pressure.[49] To mitigate risks at greater depths, enriched air nitrox mixtures with 22-40% oxygen reduce nitrogen content and extend no-decompression limits, while heliox, a helium-oxygen blend, replaces nitrogen to lower narcotic effects and gas density.[49] These mixtures are selected based on maximum operating depths to keep partial pressures within safe ranges, with gas analysis required before use.[48] Nitrogen narcosis, often called "rapture of the deep," arises from elevated partial pressures of inert gases like nitrogen, impairing central nervous system function when the nitrogen partial pressure exceeds about 3 atmospheres absolute, typically noticeable at depths of 30-40 meters on air.[50] Symptoms include euphoria, slowed reaction times, impaired judgment, and reduced manual dexterity, resembling mild alcohol intoxication, with severity increasing with depth; for instance, air dives are generally limited to 30-50 meters to avoid significant impairment.[50] Prevention involves depth restrictions or substituting helium in mixtures like heliox, which has lower narcotic potency.[49] Oxygen toxicity poses risks from high oxygen partial pressures, with central nervous system effects such as convulsions occurring above 1.6 atmospheres absolute, while pulmonary toxicity, involving lung irritation, develops above 0.5 atmospheres absolute during prolonged exposure.[48] NOAA standards limit oxygen partial pressure to 1.4 atmospheres absolute during the working phase of dives and 1.6 atmospheres absolute during decompression to minimize these hazards, tracked via oxygen toxicity units for cumulative exposure.[48] For nitrox, this determines the maximum operating depth, such as 34 meters for 32% oxygen to stay below 1.4 atmospheres absolute.[49] Decompression obligations stem from elevated partial pressures of inert gases dissolving into tissues per Henry's law, where solubility increases with pressure; rapid ascent then expands these gases per Boyle's law, forming bubbles that cause decompression sickness if not managed through staged stops.[51] Dive tables and computers calculate required stops based on tissue gas loading from partial pressures, with helium mixtures requiring longer decompression due to slower off-gassing.[48] Early diving incidents highlighted these risks and spurred gas management protocols; for example, during the 1939 USS Squalus salvage at 73 meters, U.S. Navy divers experienced severe narcosis on air, prompting the adoption of heliox mixtures for deep operations.[52] Similarly, a 1996 fatality involving oxygen toxicity at 47 meters on a 50% nitrox mix, where the oxygen partial pressure reached 2.9 atmospheres absolute causing a seizure and drowning, underscored the need for rigorous partial pressure calculations, full-face masks, and adherence to exposure limits in technical diving.[53] These events contributed to standardized training and guidelines from organizations like NOAA.[48]Medicine and Physiology
In respiratory physiology, partial pressure plays a central role in gas exchange between the alveoli and blood, driving the diffusion of oxygen and carbon dioxide across the alveolar-capillary membrane. The partial pressure of oxygen in the alveoli (P_A O_2) is estimated using the alveolar gas equation, which accounts for inspired oxygen and the respiratory exchange of gases:P_{A}O_2 \approx P_{I}O_2 - \frac{P_a CO_2}{R}
where P_{I}O_2 is the partial pressure of inspired oxygen (typically around 150 mmHg at sea level on room air), P_a CO_2 is the arterial partial pressure of carbon dioxide (approximately 40 mmHg), and R is the respiratory quotient, the ratio of carbon dioxide production to oxygen consumption, which is normally about 0.8 for a mixed diet. Under normal conditions, this yields an alveolar P_A O_2 of about 100 mmHg and P_A CO_2 of about 40 mmHg, maintaining efficient oxygenation.[54][55][56] In arterial blood, the partial pressure of oxygen (PaO_2) ranges from 75 to 100 mmHg, reflecting near-equilibration with alveolar gas, while in venous blood, PvO_2 is approximately 40 mmHg due to tissue oxygen extraction. Hemoglobin saturation with oxygen is governed by the oxygen-hemoglobin dissociation curve, a sigmoidal relationship between partial pressure and saturation percentage; at PaO_2 levels of 75-100 mmHg, arterial hemoglobin is typically 95-98% saturated, facilitating oxygen delivery to tissues, whereas the lower PvO_2 of 40 mmHg corresponds to about 75% saturation, allowing unloading in capillaries. This curve's shape ensures efficient loading in the lungs and release in tissues, with partial pressure gradients as the primary driver.[57][58][59] These gradients underpin pulmonary diffusion, as described by Fick's law, where the flux of gas across the membrane is directly proportional to the partial pressure difference (ΔP) between alveoli and blood, as well as the membrane's surface area, thickness, and gas solubility: flux ∝ ΔP × (area / thickness) × solubility. In healthy lungs, the alveolar-arterial oxygen gradient (A-a gradient) is small (5-15 mmHg), ensuring PaO_2 closely matches P_A O_2; disruptions widen this gradient, impairing oxygenation. Clinically, partial pressures are assessed via arterial blood gas (ABG) analysis, which directly measures PaO_2 and PaCO_2, or non-invasively by pulse oximetry, estimating arterial saturation (SpO_2) from light absorption in peripheral capillaries; hypoxemia is diagnosed when PaO_2 falls below 60 mmHg, often prompting interventions like supplemental oxygen.[60]/07%3A_Fundamentals_of_Gas_Exchange/7.02%3A_Fick%27s_law_of_diffusion)[57] Disorders involving low partial pressures manifest as hypoxia, with hypoxic (or hypoxemic) hypoxia arising from reduced PaO_2 due to ventilation-perfusion mismatches, diffusion impairments, or hypoventilation, leading to inadequate oxygen diffusion despite normal hemoglobin levels. In contrast, anemic hypoxia involves diminished oxygen-carrying capacity from low hemoglobin, often preserving normal PaO_2 but effectively reducing tissue oxygen availability as if partial pressures were lower; both types underscore partial pressure's role in clinical assessment and management.[61][62]