Fact-checked by Grok 2 weeks ago

Enthalpy of vaporization

The of vaporization (ΔH_vap), also known as the heat of vaporization or of vaporization, is the change in associated with the from liquid to gas at constant and , representing the required to overcome intermolecular forces and convert a unit mass or of substance from liquid to vapor without altering its . This thermodynamic property is typically expressed in units of joules per (J/kg) for specific enthalpy or kilojoules per (kJ/mol) for molar enthalpy, and it varies with , decreasing as the is approached due to weakening intermolecular attractions near the critical point. Enthalpy of vaporization is a fundamental thermophysical parameter in , essential for understanding equilibria governed by the , where for a pure substance at vapor-liquid , fixes the and vice versa. It is measured experimentally through , where the heat input required to vaporize a known of at its is quantified, often under controlled constant-pressure conditions to directly yield ΔH_vap. In practical applications, this property influences processes like , , and steam power generation, as it quantifies the of changes in systems. For , a substance with an exceptionally high ΔH_vap of approximately 40.65 /mol at 100°C, this value plays a critical role in moderating Earth's by absorbing significant heat during from oceans and land surfaces, thereby cooling environments and facilitating the water cycle's regulation of global temperatures. Across substances, ΔH_vap correlates with molecular —higher for those with strong bonding like (2257 J/g) compared to nonpolar liquids like (about 328 J/g)—highlighting its role in predicting and behavior in and .

Fundamentals

Definition

The , denoted as \Delta H_{\text{vap}}, is defined as the change in per (or per ) of a substance when it undergoes a from to gas at and . This quantifies the absorbed during , where the substance's remains despite the input of heat, as the is utilized to disrupt the molecular arrangement in the . During this process, \Delta H_{\text{vap}} represents the required to overcome the intermolecular forces that hold the molecules together, allowing them to separate and form the more disordered gaseous state. Unlike other changes such as or , the of vaporization specifically pertains to the liquid-to-vapor transition, typically evaluated at the substance's under standard pressure or at a specified and pressure condition. Mathematically, the enthalpy of vaporization is expressed as \Delta H_{\text{vap}} = H_{\text{gas}} - H_{\text{liquid}}, where H_{\text{gas}} and H_{\text{liquid}} are the molar enthalpies of the substance in the gaseous and liquid states, respectively. This difference arises because the gaseous phase has higher due to the greater separation of molecules, necessitating an endothermic process to achieve the transition.

Thermodynamic background

The enthalpy of vaporization, denoted as \Delta H_\text{vap}, is fundamentally linked to of , which states that the change in \Delta U of a equals the added q minus the work done by the w. For a at constant , the change is defined as \Delta H = \Delta U + P \Delta V, and since no non-expansion work is typically involved in reversible , \Delta H_\text{vap} = q_p, the absorbed at constant to convert one of to vapor without a change. This relation holds because the work term P \Delta V accounts for the volume expansion during the , making \Delta H the appropriate for constant-pressure processes like in an open . A key thermodynamic derivation connects \Delta H_\text{vap} to the change \Delta G via the equation \Delta G = \Delta H - T \Delta S, where T is the temperature and \Delta S is the change. At the liquid-vapor equilibrium, such as the , the process is reversible and \Delta G = 0, leading directly to \Delta H_\text{vap} = T \Delta S_\text{vap}. This equality highlights that the enthalpy of vaporization quantifies the energy required to overcome intermolecular forces in the liquid, balanced by the increased in the vapor phase, with \Delta S_\text{vap} reflecting the from the . In phase diagrams, \Delta H_\text{vap} plays a central role along the liquid-vapor coexistence curve, where the Clapeyron equation governs the of the boundary: \frac{dP}{dT} = \frac{\Delta H_\text{vap}}{T \Delta V_\text{vap}}, with \Delta V_\text{vap} being the change upon . This equation demonstrates how the enthalpy change determines the of the coexistence line, influencing the pressure-temperature conditions for ; a larger \Delta H_\text{vap} results in a steeper , reflecting greater barriers to . These derivations typically assume ideal behavior for the vapor phase, treating it as having negligible intermolecular attractions and a volume dominated by molecular motion, which simplifies \Delta V_\text{vap} \approx V_\text{gas} = RT/P. For real gases, deviations arise at high pressures or low temperatures due to finite molecular volumes and attractive forces, requiring corrections like the to account for non-ideal effects on \Delta H_\text{vap}.

Enthalpy of condensation

The of condensation, denoted as \Delta H_\text{cond}, represents the change in during the phase transition from gas to at constant , where heat is released as molecules form stronger intermolecular bonds in the state. This is exothermic, with the magnitude of \Delta H_\text{cond} quantifying the liberated per or unit mass of substance condensed. Thermodynamically, the enthalpy of condensation is the direct of the enthalpy of vaporization, such that \Delta H_\text{cond} = -\Delta H_\text{vap} when evaluated at the same temperature and pressure, ensuring across the reversible phase change. This equality holds because the forward and reverse processes involve identical intermolecular interactions, differing only in direction. In practical applications, the of condensation plays a key role in cooling processes, such as cycles, where the released heat during vapor-to-liquid transition must be dissipated to maintain system efficiency and enable continuous operation. For instance, in atmospheric contexts, condensation of releases this , warming surrounding air parcels and influencing patterns like cloud formation without altering the overall energy balance of the system.

Connection to latent heat

The enthalpy of vaporization, denoted as \Delta H_{\text{vap}}, represents the of vaporization, which quantifies the required to convert one of a substance from to gas at and . This is related to the specific of vaporization, L_{\text{vap}} (often simply called the of vaporization), by the L_{\text{vap}} = \Delta H_{\text{vap}} / [M](/page/M), where [M](/page/M) is the of the substance in /; this allows the per to be derived directly from the per-mole value. The concept of latent heat originated in the 18th century, coined by Scottish chemist during his lectures at the around 1762, where he demonstrated that is absorbed without a change during phase transitions such as the of . Black's experiments showed that requires a substantial amount of input beyond what raises its , attributing this to a "latent" or hidden form of that becomes manifest only upon phase change, with providing a prominent example alongside . A key distinction lies in their units: latent heat is typically expressed per unit mass (e.g., J/kg), making it suitable for engineering calculations involving bulk quantities, whereas the enthalpy of vaporization uses per-mole units (e.g., J/ or kJ/), aligning with thermodynamic analyses in that emphasize molecular-scale processes. This unit difference facilitates conversions via but highlights their complementary roles in different scientific contexts. In , values, including those for , are essential for calculating the total energy involved in change processes, such as determining the absorbed or released when a boils in a controlled without variation until the completes. For instance, Q = m L_{\text{vap}} (where Q is and m is ) enables precise energy balance in experiments tracking changes, aiding applications like cooling systems where or manages thermal loads. The of equals the negative of L_{\text{vap}}, reflecting the exothermic reverse process.

Measurement and units

Units

The enthalpy of vaporization, denoted as ΔH_vap, is typically expressed using units, with the given in joules per (J mol⁻¹) or kilojoules per (kJ mol⁻¹), and the specific enthalpy (per unit mass) in joules per kilogram (J kg⁻¹) or joules per gram (J g⁻¹). These units align with the () for energy and , ensuring consistency in thermodynamic reporting. Historically in chemical literature, the kilocalorie per (kcal ⁻¹) has been used as an alternative unit for molar enthalpy of vaporization, particularly in older studies before widespread adoption. In applications, especially involving and , the British thermal unit per pound (BTU lb⁻¹) serves as a common specific unit. Conversions between these units are standardized; for example, 1 kcal ⁻¹ equals exactly 4.184 kJ ⁻¹, based on the thermochemical definition of the . Similarly, 1 BTU lb⁻¹ ≈ 2.326 kJ kg⁻¹, though units are preferred for international . The International Union of Pure and Applied Chemistry (IUPAC) recommends reporting enthalpy of vaporization values under standard conditions of 298.15 K and 1 (10⁵ ) pressure, denoted with a superscript ° (e.g., ΔH_vap°), to facilitate comparison across studies, unless specified otherwise for the substance's phase behavior. This convention applies to the units described, promoting uniformity in thermodynamic data compilation.

Experimental methods

The enthalpy of vaporization, \Delta H_{\text{vap}}, is typically measured through direct calorimetric techniques that quantify the absorbed during the from to vapor at constant and . In these methods, a known quantity of is vaporized, and the energy input required to achieve complete is recorded, often using instruments designed to minimize losses and ensure isothermal conditions. Static calorimeters, sometimes referred to as batch or sealed-vessel types, involve enclosing the liquid sample in a vaporization chamber where is supplied electrically while maintaining between the and its vapor . Flow calorimeters, on the other hand, enable continuous operation by pumping through a heated section where it vaporizes, allowing for steady-state measurements of and vapor production rates, which is particularly useful for volatile or reactive substances. Indirect determination of \Delta H_{\text{vap}} relies on vapor pressure measurements over a range of temperatures, followed by application of the Clausius-Clapeyron equation. This equation derives from the thermodynamic relation for phase equilibrium, where the slope of the natural logarithm of vapor pressure versus inverse temperature yields the enthalpy change: \frac{d \ln P}{dT} = \frac{\Delta H_{\text{vap}}}{RT^2}, with P as , R the , and T the absolute ; integrating this form over a temperature range provides \Delta H_{\text{vap}} assuming it is approximately constant or using second-order corrections for temperature dependence. Vapor pressures are measured using techniques such as manometry, effusion cells, or dynamic methods like , ensuring the system reaches at each temperature point. Modern instrumental approaches have enhanced precision and automation in these measurements. (DSC) detects the endothermic peak associated with vaporization by comparing heat flows to a reference under controlled heating, allowing rapid determination of \Delta H_{\text{vap}} from the integrated area of the transition peak, often coupled with (TGA) to confirm mass loss due to evaporation. Adiabatic calorimetry, which maintains the sample thermally isolated while supplying precise heat pulses, provides high-accuracy values by minimizing external heat exchanges, particularly for low-volatility liquids where is critical. These methods are widely adopted for their to small samples and ability to operate under varying pressures. Experimental challenges include managing superheated vapors, which can lead to incomplete equilibrium or erroneous inputs if the vapor is not allowed to re-equilibrate with residual liquid. Ensuring true requires careful control of temperature gradients and vapor withdrawal rates, as deviations can introduce systematic errors in both calorimetric and data; corrections for these effects, such as accounting for in the vapor , are essential for accuracy.

Influencing factors

Temperature and pressure dependence

The enthalpy of vaporization, \Delta H_\text{vap}, exhibits a pronounced dependence on , typically decreasing as temperature increases. This variation arises from the difference in molar capacities between the vapor and phases, as described by Kirchhoff's law: \left( \frac{\partial \Delta H_\text{vap}}{\partial T} \right)_P = \Delta C_p = C_{p,\text{vapor}} - C_{p,\text{liquid}}, where \Delta C_p is generally negative because the of liquids exceeds that of gases for most substances. Integrating this relation, assuming constant \Delta C_p, yields an approximate linear decrease in \Delta H_\text{vap} with , though more accurate models account for the dependence of C_p. This trend reflects the diminishing energy required to overcome intermolecular forces as disrupts the structure. As temperature approaches the critical point, \Delta H_\text{vap} continuously diminishes and ultimately vanishes at the critical temperature T_c, where the liquid and vapor phases become indistinguishable, and the distinction between the two states ceases to exist. This critical behavior is universal across substances and stems from the coalescence of the saturation curve in the , where the required for approaches zero. Near T_c, the rate of decrease accelerates due to rapid changes in and . The dependence of \Delta H_\text{vap} on is minor under typical conditions at low pressures, owing to the low of liquids, which limits the work term's contribution to the change. However, at elevated pressures approaching the critical , becomes more substantial as the volume difference between phases narrows, amplifying deviations from ideality. The Poynting correction, which adjusts for pressure-induced changes in liquid via \ln \left( \frac{f_l}{P^\circ} \right) = \int_{P^\circ}^P \frac{V_l}{RT} \, dP, indirectly influences \Delta H_\text{vap} through integration along the saturation line, though its impact remains small except near critical conditions. For practical estimations across temperatures at saturation pressures, empirical relations like Watson's equation are employed: \Delta H_\text{vap}(T_2) = \Delta H_\text{vap}(T_1) \left( \frac{T_c - T_2}{T_c - T_1} \right)^{0.38}, derived from corresponding-states principles and validated for hydrocarbons and similar compounds, providing reliable predictions without direct measurement.

Effects in electrolyte solutions

In electrolyte solutions, the enthalpy of vaporization of the is modified by ion-solvent interactions, which alter the local structure and intermolecular forces within the liquid phase. These interactions lead to a salting-out , where ions compete for shells, effectively strengthening solvent-solvent bonds in the bulk but resulting in a slight overall decrease in the partial molar enthalpy of vaporization compared to the pure at the same . This reduction arises because the partial molar enthalpy of the in the solution is marginally higher than in the pure state, increasing the energy baseline for phase change but yielding a net lower differential enthalpy due to the of mixing. Theoretical models, such as the Debye-Hückel theory and its extensions (e.g., specific ion interaction models), describe the ionic contributions to the excess enthalpy of the solution. These frameworks predict the temperature dependence of the solvent's , allowing derivation of the partial molar enthalpy of vaporization via the integrated form of the Gibbs-Helmholtz equation or adapted Clausius-Clapeyron relations. For instance, the excess partial molar enthalpy term from electrostatic ion pairing and hydration effects typically contributes a small negative adjustment to the vaporization enthalpy in dilute solutions. Similar trends hold for , where the isobaric evaporation enthalpy at typical (35 g/kg) and 25°C is about 2439 J/g, compared to 2444 J/g for pure . These effects are particularly relevant in processes, where the reduced enthalpy of vaporization in saline solutions slightly lowers the input for the evaporation step, though this is often offset by and other factors in separation. In broader thermodynamics, understanding these modifications aids in modeling phase equilibria and energy balances for electrolyte-containing systems.

Practical values and applications

Selected values for elements

The enthalpy of vaporization (ΔH_vap) for pure is commonly reported at their normal boiling points under standard (1 atm), reflecting the required to transition from to gas phase. These values vary significantly across the periodic , influenced by bonding types, with data drawn from established thermochemical . Representative examples for metals and non-metals are tabulated below, highlighting key trends such as elevated values for metals with robust compared to those in alkali metals or non-metals with weaker intermolecular forces.

Metals

ElementBoiling Point (K)ΔH_vap (kJ/mol)
1615136
Sodium115697.4
103277
Aluminum2792290
Iron3134340
2868305
Mercury63059
Values for metals like iron and copper exceed 300 kJ/mol, underscoring the role of strong metallic bonds in requiring substantial for .

Non-metals

ElementBoiling Point (K)ΔH_vap (kJ/mol)
(H₂)200.9
(N₂)775.6
Oxygen (O₂)906.8
(F₂)856.6
(Cl₂)23920
(S₈)717.7545
Non-metals generally display lower ΔH_vap values, as seen in diatomic gases like oxygen, due to van der Waals forces rather than extended structures.

Selected values for compounds

The enthalpy of vaporization (ΔH_vap) for selected molecular compounds varies depending on molecular structure, intermolecular forces, and , typically measured at the under standard conditions. For inorganic compounds like , hydrogen bonding contributes to a relatively high value per , while for organic compounds such as and , values reflect a balance between van der Waals forces and . These data are drawn from experimentally determined thermochemical tables and are essential for understanding phase transitions in chemical processes.
CompoundFormulaBoiling Point (K)ΔH_vap (kJ/mol)Source
(inorganic)NH₃24023.35NIST Chemistry WebBook
(inorganic)H₂SO₄61056CRC Handbook of Chemistry and Physics (95th ed.) via
(organic)C₂H₅OH35138.56NIST Chemistry WebBook (Steele et al., 1997)
(organic)C₆H₆35330.72NIST Chemistry WebBook
Values are reported at or near the normal (1 atm ) and may exhibit slight variability due to measurement techniques or isotopic effects; for complex molecules like , approximations account for potential decomposition during vaporization. Standard conditions often assume ideal behavior, but real values can differ in solutions where ion-solvent interactions modify the effective .

Role in industrial processes

In industrial processes, such as those used in , the of vaporization (ΔH_vap) plays a in energy balance calculations to determine the required for separating crude into fractions like and . Engineers rely on ΔH_vap values to optimize column designs and predict vapor-liquid equilibria, ensuring efficient separation while minimizing energy consumption in multi-stage units. For instance, in atmospheric and towers, the associated with vaporizing hydrocarbons directly influences the duty and overall thermal efficiency of the . Similarly, in plants employing thermal methods like multi-stage flash (MSF) evaporation, ΔH_vap is essential for heat balance equations that account for the energy needed to vaporize into pure distillate. The gain output ratio (), a key performance metric, measures how effectively the from condensing vapor is to preheat incoming , thereby reducing overall energy input by up to several times the minimum theoretical requirement. This recycling of ΔH_vap is vital for large-scale operations, such as those producing millions of cubic meters of freshwater daily, where directly impacts operational costs and . In vapor compression refrigeration cycles, ΔH_vap governs the refrigerant's capacity to absorb heat during at low temperatures, directly affecting the (COP) and system efficiency. The process involves the refrigerant undergoing phase change in the , where it extracts from the cooled space, making high ΔH_vap values desirable for compact and energy-efficient designs in applications like and . Modern systems, including those using natural refrigerants, optimize ΔH_vap alongside other thermodynamic properties to achieve higher COPs under varying loads. Chemical process simulations, such as those performed in Aspen Plus software, integrate ΔH_vap data to model unit operations like , , and heat exchangers with high accuracy. Property methods in Aspen Plus, including those based on corresponding-states correlations, estimate or use experimental ΔH_vap to compute enthalpies and predict phase behavior in complex mixtures, enabling engineers to simulate and optimize entire plants before . This incorporation ensures reliable energy balances and process viability assessments in industries from to pharmaceuticals. Environmentally, ΔH_vap underpins evaporative cooling systems, where the high of enables efficient heat dissipation without mechanical , as seen in cooling towers for power plants and data centers. By absorbing significant during , these systems reduce ambient temperatures and conserve when designed with proper and controls. In climate modeling of the global , ΔH_vap facilitates the representation of transport, which drives and patterns, with models showing that variations in rates amplify warming effects by increasing atmospheric moisture. For example, ethanol's ΔH_vap is considered in production processes to evaluate energy recovery during .

References

  1. [1]
    Heat of Vaporization - Chemistry LibreTexts
    Jan 29, 2023 · The Heat (or Enthalpy) of Vaporization is the quantity of heat that must be absorbed if a certain quantity of liquid is vaporized at a ...Example 1 · Solution · The Enthalpy of Condensation · Example 2
  2. [2]
    Heat (Enthalpy) of Vaporization: Definition, Formula & Problems
    Oct 29, 2022 · The heat of vaporization is the enthalpy change when a unit mass of a substance changes its state from liquid to gas at a constant temperature and pressure.
  3. [3]
    [PDF] CHAPTER 6 Calorimetric Determination of Enthalpies of Vaporization
    The enthalpy of vaporization describes equilibrium between the liquid and gas phases. According to the Gibbs phase rule, only one degree of freedom (temperature ...
  4. [4]
    Equation Descriptions - Thermodynamics Research Center
    Purpose: Provide definitions of equations and equation parameters used for output of evaluated data. ... Enthalpy of Vaporization (Liquid/Gas). DEFAULT: HVP ...
  5. [5]
    [PDF] Development of standard measurement techniques and ... - GovInfo
    for heat capacity and for heat of vaporization. These techniques were to be ... The enthalpy of vaporization is by definition the enthalpy required to.
  6. [6]
    [PDF] Enthalpies of Vaporization of Organic and Organometallic ... - UMSL
    Vaporization enthalpies are important thermodynamic properties of the condensed phase. Vaporization enthalpies are used frequently in adjusting enthalpies ...
  7. [7]
    14.1 Unique Properties of Water - Maricopa Open Digital Press
    Water also has a high heat of vaporization, which is the amount of energy required to change one gram of a liquid substance to a gas. A considerable amount of ...
  8. [8]
    5.2 Enthalpy
    Enthalpy of vaporization (ΔH vap)The enthalpy change that accompanies the vaporization of 1 mol of a substance.: The enthalpy change that accompanies the ...
  9. [9]
    Heat and Enthalpy - Chemistry 301
    The heat required to convert all the liquid to a gas is called the enthalpy of vaporization, Δ H v a p o r i z a t i o n . You can see the heat required to ...
  10. [10]
    [PDF] Course Introduction
    intermolecular forces to convert a solid to a liquid. Heat (Enthalpy) of Vaporization (DH vap. ): The amount of energy required to overcome enough.
  11. [11]
    CHM1045 Enthalpy Lecture - FSU Chemistry & Biochemistry
    You may note that the units on the Enthalpy value are only shown as kJ and not kJ/mol in the reaction. This lack of the per mole unit is fairly common but the ...Missing: mass | Show results with:mass
  12. [12]
    [PDF] Thinking Like a Chemist UNIT 5 DAY 1 - Vanden Bout / LaBrake 302
    Jan 14, 2013 · Which has a higher Enthalpy? A. liquid water. B. gaseous water. C ... ΔHvaporization = Hgas - Hliquid > 0 remember: positive change in ...
  13. [13]
    [PDF] Lecture 4 6/28/10 A. Enthalpy - University of Washington
    Jun 28, 2025 · • Molar enthalpy of vaporization ∆Hvap: The heat that must be transferred at constant ... constant pressure, so the heat equals ∆H, i.e. qp=∆H. (. ).Missing: q_p | Show results with:q_p
  14. [14]
  15. [15]
    A note on integrating the Clapeyron equation without neglecting the ...
    May 15, 2023 · The derivation usually assumes that the vapor behaves as an ideal gas ... Since the enthalpy of an ideal gas depends only on temperature ...
  16. [16]
    Physical Equilibria - Chemistry 302
    Thus we see that temperature dependence of the vapor pressure depends on the enthalpy of vaporization. Remember the enthalpy term is the energy term. This is ...
  17. [17]
    2.972 How A Compression Refrigeration System Works - MIT
    Heat is given off as the temperature drops to condensation temperature. Then, more heat (specifically the latent heat of condensation) is released as the ...
  18. [18]
    Condensation and the Water Cycle | U.S. Geological Survey
    As condensation occurs and liquid water forms from the vapor, the water molecules become more organized, and heat is released into the atmosphere as a result.<|control11|><|separator|>
  19. [19]
    Unit of Heat of Vaporization — Lesson 1 - Ansys Customer Center
    The heat of vaporization is usually given in J/kg. The other SI ... Standard state enthalpy in [J/kg]* Molar Mass [kg/kgmol] = Latent Heat in [J/kgmol].
  20. [20]
    Joseph Black and Latent Heat - American Physical Society
    Having established the existence of latent heat in the melting of ice, Black turned to the vaporization of water.
  21. [21]
    University of Glasgow - Schools - School of Chemistry - Joseph Black
    He did similar work establishing the idea of latent heats of vaporization, leading to the general concept of heat capacity or specific heat. These early steps ...
  22. [22]
    11.3 Phase Change and Latent Heat - Physics | OpenStax
    Mar 26, 2020 · Explain changes in heat during changes of state, and describe latent heats of fusion and vaporization ... and both have standard units of J/kg.
  23. [23]
    RECOMMENDED REFERENCE MATERIALS FOR THE ... - iupac
    Physical property: Enthalpy of vaporization, AH1. Units: J mol -' or kJ mol '(molar enthalpy of vaporization). J kg ' or J g'(specific enthalpy of vaporization).
  24. [24]
    [PDF] Quantities, Units and Symbols in Physical Chemistry - IUPAC
    The standard chemical potential of substance B at temperature T,µ B. ◦−(T), is the value of the chemical potential under standard conditions, specified as ...
  25. [25]
    Water Properties: Vaporization Heat vs. Temperature - Charts and ...
    Online calculator, figures and tables showing heat of vaporization of water, at temperatures from 0 - 370 °C (32 - 700 °F) - SI and Imperial units.
  26. [26]
    [PDF] Thermodynamics conversion factors (pdf)
    1 kJ/kg.°C 1 kJ/kg.K = 1 J/g.°C. *Exact conversion factor between metric and English units. = = 0.062428 lbm/ft³. 1 kJ = 0.94782 Btu. == 1 Btu = 1.055056 kJ. = ...
  27. [27]
    Flow calorimeter for measuring enthalpy changes during the partial ...
    A new flow calorimeter for measuring enthalpy changes during the partial vaporization of mixtures is described. The calorimeter has been especially designed ...Missing: heat | Show results with:heat
  28. [28]
    Density, Vapor Pressure, Enthalpy of Vaporization, and Heat ...
    Mar 20, 2024 · Furtherly, the Clausius–Clapeyron relation and the Othmer method were adopted to estimate the enthalpy of vaporization of MS at the normal ...
  29. [29]
    [PDF] Vapor pressure equation for water in the range 0 to 100 deg C
    Apr 24, 2025 · We integrated the Clausius-Clapeyron equation using the accurate calorimetric data of Osborne,. Stimson, and Ginnings and the Goff and Gratch ...
  30. [30]
    [PDF] Heat of Vaporization by Measurement Using DSC/TGA ... - NREL
    Heat of vaporization (HOV) can be measured using DSC/TGA or estimated using detailed hydrocarbon analysis, but the Clausius-Clapeyron equation is problematic ...
  31. [31]
    [PDF] Measurements of Heat of Vaporization and Heat Capac
    In a vaporization experiment in a calorimeter where the liquid is in equilibrium with its vapor, heat is supplied not only to vaporize the material withdrawn ...
  32. [32]
    Universal behavior of the enthalpy of vaporization: an empirical ...
    The enthalpy of vaporization for any substance is zero at the critical point and has a maximum value at the triple point. It is a bounded function like all the ...
  33. [33]
    [PDF] Thermodynamic Properties of Aqueous Sodium Chloride Solutions
    Oct 15, 2009 · Tables of values for various thermodynamic properties are presented in the appendix. Kcx WOld"., activity coefficient5, aquCOUl5 NilC1so1utions; ...
  34. [34]
    [PDF] gsw_latentheat_evap_CT(SA,CT) - TEOS-10
    Here, we define the latent heat of evaporation as the enthalpy increase per infinitesimal mass of evaporated water of a composite system consisting of seawater ...
  35. [35]
    The system H2O–NaCl. Part II: Correlations for molar volume ...
    A set of correlations for the volumetric properties and enthalpies of phases in the system H2O–NaCl as a function of temperature, pressure, and composition ...
  36. [36]
    Periodicity » Enthalpy of vaporization » - WebElements Periodic Table
    When the pressure of the vapour of an element in equilibrium with the liquid is 1 atmosphere, the liquid element boils. Complete conversion into vapour, ...
  37. [37]
    [PDF] NIST-JANAF THERMOCHEMICAL TABLES
    The vapor pressure of the monatomic gas reaches 1 atm at 1176.9 K and the enthalpy of vaporization to monatomic gas is 23.285 kcal-mol¨¹. 800. 28.945. 88.163.
  38. [38]
    Ammonia - the NIST WebBook
    Enthalpy of vaporization ; 22.7, 308. N/A, Zander and Thomas, 1979, Based on data from 293. to 392. K.; AC ; 23.5, 239. N/A, Overstreet and Giauque, 1937, Based ...Phase change data · References
  39. [39]
    Sulfuric Acid | H2SO4 | CID 1118 - PubChem
    17 Heat of Vaporization. 56 kJ/mole. Environment Canada; Tech Info for Problem Spills: Sulfuric acid and Oleum (Draft) p.6 (1984). Hazardous Substances Data ...
  40. [40]
    Ethanol - the NIST WebBook
    Enthalpy of vaporization ; 14.2, 503. C · Vine and Wormald, 1989, AC ; 40.9, 320. C · Dong, Lin, et al., 1988, AC.
  41. [41]
  42. [42]
    [PDF] Thermal properties of petroleum products - GovInfo
    ... heat of vaporization of petroleum oils, given in Table 15, were calculated from the equation. Z=|(110.9-0.09f). (6) in which Z = latent heat of vaporization ...
  43. [43]
    [PDF] Thermal Desalination Process
    Furthermore, assume that the specific heat capacity of the liquid CP and latent heat of vaporization λ do not vary in the range of operation. Heat balance of ...
  44. [44]
    [PDF] Entropy Generation Analysis of Desalination Technologies
    Sep 30, 2011 · In essence, GOR is a measure of how many times the latent heat of vaporization is captured in the condensation of pure water vapor and reused in ...
  45. [45]
    Vapor Compression Cycle: A State-of-the-Art Review on ... - MDPI
    This study presents the latest developments in the vapor compression cycle and natural refrigerants, focusing on water as a refrigerant.Vapor Compression Cycle: A... · 2. Refrigeration Cycle... · 3. Refrigerants Development...
  46. [46]
    Thermodynamics University Modules - AspenTech Support
    This module will review the concept of enthalpy. We find the heat of vaporization of Benzene using Pure Analysis to demonstrate how to set up and interpret ...
  47. [47]
    Evaporative Cooling Systems: How and Why They Work
    To evaporate water, heat (energy) is required. The heat comes from whatever object the water is in contact with as it evaporates. In our cool cell situation, ...
  48. [48]
    WATER VAPOR AND THE DYNAMICS OF CLIMATE CHANGES
    Jul 2, 2010 · At the reference simulation, the water vapor cycling rate decreases with global-mean surface temperature at 3.7% K−1, the difference ...
  49. [49]
    The Water Cycle and Climate Change - NASA Earth Observatory
    Oct 1, 2010 · For example, as the lower atmosphere becomes warmer, evaporation rates will increase, resulting in an increase in the amount of moisture ...Missing: enthalpy | Show results with:enthalpy<|control11|><|separator|>