Fact-checked by Grok 2 weeks ago

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. 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. 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. The partial pressure of a gas, denoted as P_i, is quantitatively determined by multiplying the pressure P of the by the 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. This relationship holds for ideal gases and is fundamental in gas , enabling calculations of gas compositions in reactions and mixtures. In practical terms, partial s are crucial for understanding gas behavior in diverse fields, including , where they describe the contributions of (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. Beyond chemistry, partial pressures play a vital role in , particularly in respiratory , where the partial pressure of oxygen (P_{O_2}) in alveolar air drives into the bloodstream, and elevated P_{CO_2} in tissues promotes release. In clinical settings, monitoring arterial partial pressures via blood gas analysis assesses oxygenation, , and acid-base balance, guiding treatments for conditions like . Additionally, the concept is essential in applications such as , where increased total pressure at depth raises partial pressures of inert gases, risking , and in like and efficiency optimization.

Basic Principles

Definition

Partial pressure is defined as the hypothetical pressure exerted by an individual gas component in a if that gas alone occupied the entire of the at the same as the . This measure assumes that the gases in the do not chemically interact and behave independently, allowing each to contribute proportionally to the total based on its . 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. 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. This 1801 insight, later formalized as Dalton's law of partial pressures, provided a key empirical basis for understanding gaseous systems. To intuitively grasp partial pressure, consider a of gases as akin to multiple individuals in a , each generating independently that collectively determines the total noise level, without one altering the output of another. In a practical example, Earth's atmosphere at has a total of about 101 kPa, with oxygen comprising roughly 21% of dry air, yielding an oxygen partial pressure of approximately 21 kPa.

Notation and Units

In , 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 or 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 of component i and p is the total pressure. Partial pressures are quantified using the same units as measurements. The unit is the pascal (), equivalent to one per square meter (N/), which provides a standardized for precise calculations in and gas dynamics. In practice, especially in and , other units are prevalent, including the atmosphere (), , millimeter of mercury (mmHg), and ; for instance, normal is 1 , or exactly 101325 . Conversions between these units are essential: 1 = 100000 , 1 mmHg = 133.322 , and 1 ≈ 133.322 (with 1 defined as exactly 1/760 )./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 , ensuring unambiguous representation in formulas like those for constants or . The torr unit originated from the experiments of Italian physicist in 1643, who developed the mercury barometer to measure as the height of a mercury column, with one 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.

Dalton's Law

Dalton's law of partial pressures states that in a of non-reacting gases, the total 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. The partial pressure of a component gas P_i is given by P_i = x_i P_{\text{total}}, where x_i is the 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. This law can be derived from the kinetic molecular theory of gases, which posits that gas arises from the random collisions of with the container walls. In a , of each gas collide independently with the walls, and the total is the sum of contributions from each , proportional to their respective number densities (moles per unit volume) since the average per is the same for all gases at a given ./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 ; it holds well under conditions of low and high temperature where these assumptions are valid. formulated the law in 1801 based on his meteorological observations, particularly experiments measuring the of in air, where he found that the vapor's remained unchanged regardless of the presence of dry air, confirming the additive nature of pressures in mixtures. In real gases, deviations from occur at high pressures due to intermolecular attractions and the finite volume of molecules, which can be partially accounted for using equations like the for mixtures, though the law remains a good approximation for most practical scenarios involving dilute gases.

Gas Mixture Properties

Ideal Gas Approximations

In ideal gas mixtures, the composition is quantified using the , 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}}}. This 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. This connection stems from , which posits that the total pressure is the sum of individual partial pressures in non-interacting ideal gases. 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. 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. 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. 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.

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 . This formulation connects directly to partial pressure via the . 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 (since p_i / P = x_i). The law originated from experiments by French physicist Émile-Hilaire Amagat on the 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 mixtures. Amagat's law complements by offering a volume-additive viewpoint at fixed and , in contrast to Dalton's pressure-additive approach at fixed volume. Both describe mixtures but apply to distinct conditions, enabling consistent predictions of mixture properties from either perspective. A representative example is the mixing of 1 of and 1 of oxygen at 298 and 1 total pressure. Each gas alone would occupy 24.5 L (from the ), so the mixture volume is 49.0 L with no deviation upon ideal mixing.

Vapor Pressure

Vapor pressure refers to the partial pressure exerted by a vapor in with its liquid or solid phase in a at a given ./Physical_Properties_of_Matter/States_of_Matter/Properties_of_Liquids/Vapor_Pressure) This equilibrium arises when the rate of equals the rate of , resulting in a constant pressure attributable solely to the vapor component. In mixtures containing non-condensable gases, the represents the partial pressure of the condensable component, distinct from the contributions of other gases. The of a substance increases with due to enhanced molecular , which favors over . This 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 , T is the absolute , \Delta H_\text{vap} is the , and R is the ./Physical_Properties_of_Matter/States_of_Matter/Phase_Transitions/Clausius-Clapeyron_Equation) The equation highlights the exponential relationship between and , explaining why occurs when equals the surrounding total pressure. A representative example is , whose at 25°C is 3.17 kPa (23.8 mmHg). In atmospheric contexts, this plays a key role in humidity metrics; relative humidity (RH) is defined as the ratio of the actual partial pressure of (P_\text{vap}) to the (P_\text{sat}) at that , expressed as RH = (P_\text{vap} / P_\text{sat}) \times 100\%. For instance, at 25°C, an RH of 50% corresponds to a partial pressure of 1.585 kPa. In moist air, the atmospheric pressure (P_\text{total}) is the sum of the partial pressure of air (P_\text{dry}) and the partial pressure of (P_\text{vap}), following of partial pressures: P_\text{total} = P_\text{dry} + P_\text{vap}. This distinction is crucial, as P_\text{vap} can approach P_\text{sat} near saturation, potentially leading to if exceeded. Vapor pressure is measured using static or dynamic techniques. Static methods involve equilibrating the sample in a closed and directly gauging the , often with manometers to avoid . Dynamic methods, such as gas carrier techniques, pass a dry over the and measure the increase in or downstream. A specific static variant is the isoteniscope method, which uses a capillary tube to trap a vapor bubble and indirectly determines 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. The law originates from observations by William Henry in 1803, who found that the quantity of gas dissolved in under is proportional to the degree of compression, marking an early quantitative study of gas . 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 of the solute in the liquid; solubility-based variants express the ratio of gas volume dissolved to gas volume at standard . 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. 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 as solubility drops. Similarly, in oxygenation processes like for , 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. As a limiting law analogous to the , 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 effects. These limitations restrict its use to scenarios like sparingly soluble gases in solvents, ensuring accuracy in predictions of partitioning.

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). This formulation arises because partial pressures directly reflect the effective concentration of gases in mixtures under ideal conditions, influencing the position of . 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 , 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). This connection highlights how partial pressures scale with total pressure and temperature, affecting shifts when \Delta n \neq 0. 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. At equilibrium, \Delta G^\circ = -RT \ln K_p, linking standard free energy to partial pressure-based constants. A key example is the Haber-Bosch process for synthesis, N_2 + 3H_2 \rightleftharpoons 2NH_3, where \Delta n = -2, so increasing total pressure raises partial pressures of reactants and shifts toward products per , enhancing yield despite the exothermic nature favoring lower temperatures. Industrial conditions typically use 200–300 atm and 400–500°C with iron catalysts to balance and thermodynamics. In industrial , partial pressures guide reactor design for processes like or synthesis, where high reactant partial pressures maximize K_p-driven conversions. For equilibria, such as in gas turbines, partial pressures of oxygen and fuels determine 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. 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. 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. These mixtures are selected based on maximum operating depths to keep partial pressures within safe ranges, with gas analysis required before use. Nitrogen narcosis, often called "rapture of the deep," arises from elevated partial pressures of inert gases like , impairing function when the nitrogen partial pressure exceeds about 3 atmospheres absolute, typically noticeable at depths of 30-40 meters on air. Symptoms include , slowed reaction times, impaired judgment, and reduced manual dexterity, resembling mild , with severity increasing with depth; for instance, air dives are generally limited to 30-50 meters to avoid significant impairment. Prevention involves depth restrictions or substituting in mixtures like , which has lower narcotic potency. 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. NOAA standards limit oxygen partial pressure to 1.4 atmospheres absolute during the working phase of dives and 1.6 atmospheres absolute during to minimize these hazards, tracked via oxygen toxicity units for cumulative exposure. For , this determines the maximum operating depth, such as 34 meters for 32% oxygen to stay below 1.4 atmospheres absolute. Decompression obligations stem from elevated partial pressures of inert gases dissolving into tissues per , where solubility increases with pressure; rapid ascent then expands these gases per , forming bubbles that cause if not managed through staged stops. 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. Early diving incidents highlighted these risks and spurred gas management protocols; for example, during the USS Squalus salvage at 73 meters, U.S. divers experienced severe narcosis on air, prompting the adoption of mixtures for deep operations. Similarly, a 1996 fatality involving at 47 meters on a 50% mix, where the oxygen partial pressure reached 2.9 atmospheres absolute causing a and , underscored the need for rigorous partial pressure calculations, full-face masks, and adherence to exposure limits in . These events contributed to standardized training and guidelines from organizations like NOAA.

Medicine and Physiology

In respiratory , partial pressure plays a central role in between the alveoli and blood, driving the diffusion of oxygen and 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 on room air), P_a CO_2 is the arterial partial pressure of (approximately 40 mmHg), and R is the , 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.
In , the partial pressure of oxygen (PaO_2) ranges from 75 to 100 mmHg, reflecting near-equilibration with alveolar gas, while in , PvO_2 is approximately 40 mmHg due to oxygen extraction. saturation with oxygen is governed by the oxygen- dissociation curve, a sigmoidal between partial pressure and percentage; at PaO_2 levels of 75-100 mmHg, arterial is typically 95-98% saturated, facilitating oxygen delivery to , whereas the lower PvO_2 of 40 mmHg corresponds to about 75% , allowing unloading in capillaries. This curve's shape ensures efficient loading in the lungs and release in , with partial pressure gradients as the primary driver. These gradients underpin pulmonary diffusion, as described by Fick's law, where the of gas across the 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 () analysis, which directly measures PaO_2 and PaCO_2, or non-invasively by , estimating arterial saturation (SpO_2) from light absorption in peripheral capillaries; is diagnosed when PaO_2 falls below 60 mmHg, often prompting interventions like supplemental oxygen./07%3A_Fundamentals_of_Gas_Exchange/7.02%3A_Fick%27s_law_of_diffusion) 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 , leading to inadequate oxygen diffusion despite normal levels. In contrast, anemic hypoxia involves diminished oxygen-carrying capacity from low , 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.

Atmospheric and Environmental Uses

In Earth's atmosphere, partial pressures of individual gases are determined by their mole fractions multiplied by the total , which is approximately 101.3 kPa at under standard conditions. constitutes the largest fraction, with a partial pressure of about 78 kPa, while oxygen contributes roughly 21 kPa, and around 0.043 kPa (≈426 as of November 2025). These values reflect the composition of dry air, where partial pressure varies but is typically low at , influencing overall gas dynamics. As altitude increases, total declines exponentially, causing partial pressures of all gases to decrease proportionally, though their relative fractions remain largely constant in the . For oxygen, the partial pressure halves from its sea-level value of approximately 21 kPa to about 10.5 kPa at around 5,500 meters, due to the barometric pressure reduction following the hydrostatic equation. This gradient is critical for understanding atmospheric layering and gas availability in upper regions. Partial pressures of greenhouse gases like and play a central role in the that drives the , by absorbing and re-emitting . 's partial pressure, currently around 0.043 kPa corresponding to a global average concentration of approximately 426 as of November 2025, contributes to positive of about 2.16 W/m² since pre-industrial times, amplifying surface warming. , as the most abundant , acts as a mechanism; its partial pressure increases with , enhancing the effect by roughly doubling the warming from CO2 alone, though it is not a primary forcing agent. In air quality monitoring, the partial pressure of tropospheric is a key indicator of levels, typically ranging from 10 to 100 (ppb) by volume, equivalent to about 0.001 to 0.01 kPa at . Background concentrations in remote areas average 20-40 ppb, while urban can elevate levels to 80-100 ppb, contributing to photochemical reactions that degrade air quality and form ground-level oxidants. Regulatory standards, such as the U.S. EPA's 70 ppb 8-hour average, target these elevated partial pressures to mitigate and environmental risks. Climate models incorporate partial pressures to simulate hydrological processes, particularly by comparing the partial pressure of to the saturation , which determines relative humidity and drives and rates. In these models, deviations from saturation influence cloud formation and moisture transport, with higher partial pressures in warmer climates intensifying precipitation efficiency while altering fluxes over oceans and land. This approach ensures accurate representation of the water cycle's response to radiative perturbations. Recent studies highlight the role of partial pressures in -atmosphere CO2 exchange, where surface levels—often exceeding 400 µatm in regions—drive net fluxes that exacerbate acidification by lowering pH. For instance, research in fjords shows spatial variability from 200 to 500 µatm, influenced by temperature and freshwater inputs, affecting carbon uptake and biological . Similarly, analyses of the Subpolar North Atlantic reveal emerging "CO2 uptake holes" where rising reduces capacity under warming, potentially weakening global sinks by 10-20% by mid-century. These dynamics underscore partial pressure gradients as predictors of acidification impacts on ecosystems.

References

  1. [1]
    The Respiratory System and Oxygen Transport - NCBI - NIH
    Dalton's law of partial pressures defines the partial pressure of a gas in a gas mixture as the pressure that the gas would exert if it occupied the total ...
  2. [2]
    9.4 Mixtures of Gases and Partial Pressures
    Dalton's law of partial pressures: total pressure of a mixture of ideal gases is equal to the sum of the partial pressures of the component gases.
  3. [3]
    State of shock: 200-year-old law about gas mixtures called into ...
    Dec 9, 2019 · And in 1802, scientist John Dalton stated that the total pressure in a non-reactive gas mixture – at constant temperature and volume – is equal ...
  4. [4]
    Gas Behavior - FSU Chemistry & Biochemistry
    Dalton's Law of Partial Pressures: The total pressure exerted by a mixture of gases is the sum of their individual partial pressures. Ptotal = P1 + P2 + P3 + ….
  5. [5]
    Partial Pressure - Chembook
    The concept of partial pressure comes from the fact that each specific gas contributes a part of the total pressure and that part is the partial pressure of ...Missing: definition | Show results with:definition
  6. [6]
    12.3 Dalton's Law of Partial Pressures
    A law that states that the total pressure exerted by a mixture of gases is the sum of the partial pressures of component gases.Missing: definition | Show results with:definition
  7. [7]
    Gas Laws and Clinical Application - StatPearls - NCBI Bookshelf
    Dalton's law of partial pressures states that, for a mixture of non-reacting gases, the sum of the partial pressure of each gas is equal to the total pressure ...
  8. [8]
    Partial Pressure of Carbon Dioxide - StatPearls - NCBI Bookshelf
    Sep 26, 2022 · The most common use of PCO2 is monitoring respiratory and acid-base status in patients receiving mechanical ventilation. Respiratory care ...<|control11|><|separator|>
  9. [9]
    What Is Partial Pressure - Salem State Vault
    Dalton's Law of Partial Pressures states that the total pressure of a mixture of gases is equal to the sum of the partial pressures of the individual gases.
  10. [10]
    Partial Pressure of Oxygen - StatPearls - NCBI Bookshelf - NIH
    At sea level, the atmospheric pressure is 760 mmHg. However, as elevation increases, atmospheric pressure decreases. For example, the atmospheric pressure is as ...Introduction · Function · Clinical Significance
  11. [11]
  12. [12]
    Partial Pressure Gradients in the Lung - EdTech Books
    For example, at sea level the total pressure is 760 mm Hg and oxygen makes up 21% of the gas, the partial pressure of oxygen in the atmosphere is 160 mm Hg ( ...
  13. [13]
    pressure (P04819) - IUPAC Gold Book
    For a mixture of gases the contribution by each constituent is called the 'partial pressure' p i = x i p , where x i is the amount fraction of the i th ...
  14. [14]
    [PDF] Quantities, Units and Symbols in Physical Chemistry - IUPAC
    it is clearly an aid to scientific communication if we all generally follow a standard notation. ... standard pressure p◦−, the standard molality m◦−,.
  15. [15]
    Evangelista Torricelli
    To recognize Torricelli's contributions, some scientists describe pressure in units of "torr," which are defined as follows. 1 torr = 1 mmHg. fig4_5.gif ...
  16. [16]
    Evangelista Torricelli and the mercury barometer - Leybold USA
    The first mercury filled glass tube was generally accepted to have been invented in 1643 by Torricelli, which became known as Torricelli's mercury barometer.
  17. [17]
  18. [18]
    Dalton's law of partial pressure (article) | Khan Academy
    Dalton's law of partial pressures states that the total pressure of a mixture of gases is the sum of the partial pressures of its components.
  19. [19]
  20. [20]
  21. [21]
    Mole Fraction and Partial Pressure of the Gas - Chemistry Steps
    The partial pressure of a gas in a mixture is equal to the product of its mole fraction and the total pressure of all gases:Missing: definition | Show results with:definition
  22. [22]
    10.6: Gas Mixtures and Partial Pressures - Chemistry LibreTexts
    Jun 24, 2024 · Use the ideal gas law to calculate the partial pressure of each gas. Then add together the partial pressures to obtain the total pressure of ...Partial Pressures · Example 10 . 6 . 1 : The Bends · Mole Fractions of Gas Mixtures
  23. [23]
    Gas Mixtures - Chemistry 301
    The partial pressure of any substance in a gas mixture is the mole fraction of that substance times the total pressure.
  24. [24]
    Mixtures of Gases; Partial Pressures
    Example: Dry air contains 78.08% nitrogen, 20.095% oxygen, and 0.93% argon. Calculate the partial pressure of each gas in a sample of dry air at 760 torr.
  25. [25]
    [PDF] Equations of State Ideal Gas
    Changes in internal energy and enthalpy of mixtures u2 − u1. = X Xi(u2 − u1)i = Z T2. T1 cv dT = cv(T2 − T1) h2 − h1. = X Xi(h2 − h1)i = Z T2. T1 cp dT ...<|control11|><|separator|>
  26. [26]
    Dalton's and Amagat's laws fail in gas mixtures with shock propagation
    Dec 6, 2019 · In 1802, John Dalton's publication in Memoirs of the Literary and Philosophical Society of Manchester formulated the law of additive (or partial) ...
  27. [27]
    [PDF] Non-Reacting Gas Mixtures
    Amagat Model (law of additive volumes). • the volume of ... For a mixture of ideal gases the pressure is the sum of the partial pressures of the individual.<|control11|><|separator|>
  28. [28]
    Vapor Pressure
    Vapor pressure is the equilibrium pressure of a vapor above a liquid or solid, the pressure exerted by the gas in equilibrium with the solid or liquid.
  29. [29]
    [PDF] Vapor pressure - IDC Technologies
    Vapor pressure is the pressure of a vapor in equilibrium with its liquid or solid phase, when molecules escape and return at equal rates.
  30. [30]
    Water - the NIST WebBook
    Water's formula is H2O, molecular weight is 18.0153, and CAS Registry Number is 7732-18-5. Other names include water vapor and distilled water.
  31. [31]
    Definition of relative humidity - Conservation Physics
    Relative humidity (RH) is the ratio of the actual partial vapour pressure of water to the saturation vapour pressure at that temperature.<|separator|>
  32. [32]
    Moist Air - Partial Pressure and Daltons Law
    The pressure in a mixture of dry air and water vapor - humid or moist air - can be estimated by using Daltons Law of partial pressures.Missing: P_vap | Show results with:P_vap
  33. [33]
    [PDF] Henry's law constants - OSTI.GOV
    Jan 3, 2021 · Henry's law states that a solute's abundance in a liquid is proportional to its gas phase abundance. The constant is a proportionality ...
  34. [34]
    (PDF) Henry's law constants (IUPAC Recommendations 2021)
    Henry's law states that the abundance of a volatile solute dissolved in a liquid is proportional to its abundance in the gas phase.
  35. [35]
    Henry's Law Constants
    IUPAC recommends eight variants of Henry's law constants (Sander et al., 2022): Here, the Henry's law solubility constant H s cp is used. It is defined as the ...
  36. [36]
    Henry's Law - StatPearls - NCBI Bookshelf - NIH
    This law defines the relationship between the partial pressure of gases overlying a solution and the gases' ability to dissolve in that solution.
  37. [37]
    [PDF] Using Henrys Law To Calculate The Solubility Of A Gas
    Dec 27, 2013 · This article delves into the specifics of Henry's Law, its mathematical formulation, and its applications in calculating the solubility of gases ...
  38. [38]
    Kp - Chemistry LibreTexts
    Jan 29, 2023 · This page explains equilibrium constants expressed in terms of partial pressures of gases, K p. It covers an explanation of the terms mole fraction and partial ...
  39. [39]
    Calculating equilibrium constant Kp using partial pressures (article)
    When the equilibrium constant is written with the gases in terms of partial pressure, the equilibrium constant is written as the symbol K p ‍ . The subscript p ...
  40. [40]
    Gas Equilibrium Constants - Chemistry LibreTexts
    Jan 29, 2023 · \(K_c\) and \(K_p\) are the equilibrium constants of gaseous mixtures. However, the difference between the two constants is that \(K_c\) is ...Example \(\PageIndex{1... · Relating Gas Equilibrium... · How the Gas Equilibrium...
  41. [41]
    Converting Between K c and K p
    Kp = Kc(RT)Dn​​ Dn is the difference in the number of moles of gases on each side of the balanced equation for the reaction. Calculate the difference in the ...
  42. [42]
    7.11 Gibbs Free Energy and Equilibrium - Chemistry LibreTexts
    Jun 5, 2019 · In this section, we explore the relationship between the free energy change of reaction (ΔG) and the instantaneous reaction quotient (Q).
  43. [43]
    Changes in free energy and the reaction quotient - Khan Academy
    Mar 30, 2016 · The relationship between Gibbs free energy, standard Gibbs free energy, and the reaction quotient Q. Predicting the direction of the reaction when we don't ...
  44. [44]
    The Haber Process - Chemistry LibreTexts
    Jan 29, 2023 · According to Le Chatelier's Principle, if you increase the pressure the system will respond by favoring the reaction which produces fewer ...Missing: partial | Show results with:partial
  45. [45]
    De-carbonising Ammonia: alternatives to the Haber Bosch process?
    The Haber Bosch process illustrates Le Chatelier's principle very well. A reversible reaction in a closed system will reach equilibrium. Changes in applied ...
  46. [46]
    [PDF] Dehydrogenation by Heterogeneous Catalysts
    The major cause of catalyst deactivation in dehydrogenation processes is carbon deposition. The high temperatures and low pressures necessary to achieve high.
  47. [47]
    None
    Below is a merged summary of the NOAA Diving Standards & Safety Manual (May 2023) based on the provided segments. To retain all information in a dense and organized format, I’ve used tables where appropriate, followed by a narrative summary for additional details. The response consolidates all topics across the segments, prioritizing the most specific and detailed information while avoiding redundancy.
  48. [48]
    Breathing Gases - Divers Alert Network
    May 1, 2017 · The widely accepted safe partial pressure limit for oxygen during the working or bottom phase of a dive is 1.4, and most training organizations ...
  49. [49]
    Nitrogen Narcosis In Diving - StatPearls - NCBI Bookshelf - NIH
    Limiting the depth of a dive is one of the least invasive. It is agreed upon that the maximum depth limit for a diver to use compressed air is 30 to 50 meters.Missing: threshold | Show results with:threshold
  50. [50]
    The pathophysiologies of diving diseases - PMC - PubMed Central
    Decompression sickness. During a dive, the increased partial pressure of an inert gas breathed causes tissues to take up greater amounts of the dissolved gas ...Effects Of Ascent: Gas... · Decompression Sickness · Fig 1
  51. [51]
    A Practical Discussion of Nitrogen Narcosis for Deep Diving -
    Jul 17, 2012 · Most divers will be able to function well in excess of the so-called 130 fsw (39.4 m) limit with even a little practice.Missing: threshold | Show results with:threshold
  52. [52]
    A diving fatality due to oxygen toxicity during a
    Sep 2, 1996 · An experienced diver drowned after a generalised seizure caused by oxygen toxicity during a 19-minute "technical" dive to a depth of 47 m.
  53. [53]
    Alveolar Gas Equation - StatPearls - NCBI Bookshelf - NIH
    The alveolar gas equation calculates alveolar oxygen partial pressure, estimating P_AO2 inside the alveoli, as it's impossible to collect gases directly.Introduction · Function · Issues of Concern · Clinical Significance
  54. [54]
    Physiology, Respiratory Quotient - StatPearls - NCBI Bookshelf - NIH
    The respiratory quotient, or the respiratory ratio (RQ), is the volume of carbon dioxide released over the volume of oxygen absorbed during respiration.
  55. [55]
    14. The Alveolar Gas Equation and Alveolar–Arterial PO2 Difference
    The alveolar gas equation estimates whole lung alveolar PO 2 as the inspired PO 2 minus the arterial PO 2 divided by the respiratory exchange ratio.
  56. [56]
    Arterial Blood Gas - StatPearls - NCBI Bookshelf
    Jan 8, 2024 · pH (7.35-7.45). PaO2 (75-100 mm Hg). PaCO2 (35-45 mm Hg).Arterial Blood Gas · Specimen Requirements And... · Clinical SignificanceMissing: PvO2 | Show results with:PvO2
  57. [57]
    Blood Gas, Venous, McLendon Clinical Laboratories | UNC Hospitals
    Dec 14, 2023 · Blood Gas, Venous ; pH, 7.32-7.43, <7.20 or >7.65 ; pCO · 40-60 mmHg, <20 or >65 mmHg ; pO · 30-55 mmHg (at Room Air), N/A ; HCO · 22-27 mmol/L · N/A.
  58. [58]
    Physiology, Oxygen Transport And Carbon Dioxide Dissociation Curve
    [4] In the pulmonary capillaries, the partial pressure of oxygen is high, allowing more oxygen molecules to bind hemoglobin until reaching the maximum ...Introduction · Mechanism · Clinical Significance
  59. [59]
    Physiology, Pulmonary Ventilation and Perfusion - StatPearls - NCBI
    According to Fick's law of diffusion, diffusion of a gas across the alveolar membrane increases with: Increased surface area of the membrane. Increased ...
  60. [60]
    Hypoxia - StatPearls - NCBI Bookshelf - NIH
    Mar 4, 2024 · Anemic hypoxia: Caused by a reduction in oxygen-carrying capacity due to low hemoglobin levels, resulting in insufficient oxygen delivery.
  61. [61]
    Acute Respiratory Failure - All There Is To Know - Pinson & Tang, LLC
    Sep 5, 2023 · The gold standard for the diagnosis of acute hypoxemic respiratory failure is an arterial pO2 on room air less than 60 mmHg measured by arterial ...
  62. [62]
    [PDF] US Standard Atmosphere, 1976
    The quantity go (= 9.80665 m/s²) repre- sents the sea-level value of the acceler-.
  63. [63]
    [PDF] COMPOSITION, CHEMISTRY, AND CLIMATE OF THE ATMOSPHERE
    Trace gases include ozone, methane (CH4), oxides of nitrogen, carbon monoxide (CO), and halocarbons. The proportions of the major dry atmospheric gases are ...
  64. [64]
    [PDF] Habitable Atmosphere OCHMO-TB-003 Rev A - Executive Summary
    Sea Level Total Air Pressure. 1 ATM=760 mmHg=14.7psia=101.3 kPa. Standard Sea-Level Atmosphere. Source: NASA HIDH. Atmosphere Measurement. • Total pressure ...
  65. [65]
    High-altitude physiology and pathophysiology - PubMed Central - NIH
    For reference, the partial pressure of oxygen at the altitude of Everest Base Camp (5,300 metres altitude) is about one-half of the sealevel value. The ...
  66. [66]
    Trends in CO 2 , CH 4 , N 2 O, SF 6 - Global Monitoring Laboratory
    Sep 5, 2025 · The global monthly mean CO2 was 425.83 ppm in June 2025, and 423.22 ppm in June 2024. Data is measured at marine surface sites.
  67. [67]
    [PDF] Anthropogenic and Natural Radiative Forcing
    As the largest contributor to the natural greenhouse effect, water vapour plays an essential role in the Earth's ... from the albedo change with the greenhouse ...
  68. [68]
    [PDF] CHAPTERS Tropospheric Ozone
    The observed correlation between ozone and alkyl nitrates suggests a natural ozone concentration of 20-30 ppb in the upper planetary boundary layer (at ...Missing: partial | Show results with:partial
  69. [69]
    Ground-level Ozone Basics | US EPA
    Mar 11, 2025 · Learn the difference between good (stratospheric) and bad (tropospheric) ozone, how bad ozone affects our air quality, health, ...Missing: ppb | Show results with:ppb
  70. [70]
    [PDF] water vapor feedback and global warming
    Oct 16, 2000 · To the extent that water vapor concentrations increase in a warmer world, the climatic effects of the other greenhouse gases will be amplified.
  71. [71]
    Unraveling ocean pCO 2 dynamics in Northwest Greenland Fjords
    Aug 11, 2025 · Our findings revealed significant spatial variability in pCO 2 , driven by differences in temperature, freshwater inputs, and biological ...
  72. [72]
    Emergence of an oceanic CO 2 uptake hole under global warming
    Apr 3, 2025 · We find a pronounced weakening of ocean CO 2 uptake in the Subpolar North Atlantic (SPNA), distinct from the global response.Missing: pCO2 | Show results with:pCO2