Sie sind auf Seite 1von 15

Liquid-vapor critical point[edit]

Overview[edit]

The liquid-vapor critical point in a pressure–temperature phase diagram is at the high-temperature extreme of the liquid–gas
phase boundary. The dotted green line shows the anomalous behavior of water.

For simplicity and clarity, the generic notion of critical point is best introduced by discussing a specific example, the
liquid-vapor critical point. This was the first critical point to be discovered, and it is still the best known and most
studied one.
The figure to the right shows the schematic PT diagram of a pure substance (as opposed to mixtures, which have
additional state variables and richer phase diagrams, discussed below). The commonly
known phases solid, liquid and vapor are separated by phase boundaries, i.e. pressure-temperature combinations
where two phases can coexist. At the triple point, all three phases can coexist. However, the liquid-vapor boundary
terminates in an endpoint at some critical temperature Tc and critical pressure pc. This is the critical point.
In water, the critical point occurs at around 647 K (374 °C or 705 °F) and 22.064 MPa (3200 psia or 218 atm).[2]
In the vicinity of the critical point, the physical properties of the liquid and the vapor change dramatically, with both
phases becoming ever more similar. For instance, liquid water under normal conditions is nearly incompressible,
has a low thermal expansion coefficient, has a high dielectric constant, and is an excellent solvent for electrolytes.
Near the critical point, all these properties change into the exact opposite: water becomes compressible,
expandable, a poor dielectric, a bad solvent for electrolytes, and prefers to mix with nonpolar gases and organic
molecules.[3]
At the critical point, only one phase exists. The heat of vaporization is zero. There is a stationary inflection point in
the constant-temperature line (critical isotherm) on a PV diagram. This means that at the critical point: [4][5][6]
The critical isotherm with the critical point K

Above the critical point there exists a state of matter that is continuously connected with (can be
transformed without phase transition into) both the liquid and the gaseous state. It is called supercritical
fluid. The common textbook knowledge that all distinction between liquid and vapor disappears beyond the
critical point has been challenged by Fisher and Widom[7] who identified a p,T-line that separates states with
different asymptotic statistical properties (Fisher-Widom line).

History[edit]

Carbon dioxide exuding fog while cooling from supercritical to critical temperature

The existence of a critical point was first discovered by Charles Cagniard de la Tour in 1822[8][9] and named
by Dmitri Mendeleev in 1860[10][11] and Thomas Andrews in 1869.[12] Cagniard showed that CO2 could be
liquefied at 31 °C at a pressure of 73 atm, but not at a slightly higher temperature, even under pressures as
high as 3,000 atm.

Theory[edit]
Solving the above condition  for the van der Waals equation, one can compute the critical point as
.
However, the van der Waals equation, based on a mean field theory, does not hold near the critical
point. In particular, it predicts wrong scaling laws.
To analyse properties of fluids near the critical point, reduced state variables are sometimes defined
relative to the critical properties[13]
.
The principle of corresponding states indicates that substances at equal reduced pressures and
temperatures have equal reduced volumes. This relationship is approximately true for many
substances, but becomes increasingly inaccurate for large values of pr.
For some gases, there is an additional correction factor, called Newton's correction, added to the
critical temperature and critical pressure calculated in this manner. These are empirically derived
values and vary with the pressure range of interest. [14]

Table of liquid–vapor critical temperature and pressure for


selected substances[edit]
See also: Critical points of the elements (data page)

Substance[15][16] Critical temperature Critical pressure (absolute)

Argon −122.4 °C (150.8 K) 48.1 atm (4,870 kPa)

Ammonia (NH3)[17] 132.4 °C (405.5 K) 111.3 atm (11,280 kPa)

R-134a 101.06 °C (374.21 K) 40.06 atm (4,059 kPa)

R-410A 72.8 °C (345.9 K) 47.08 atm (4,770 kPa)

Bromine 310.8 °C (584.0 K) 102 atm (10,300 kPa)

1,664.85 °C
Caesium 94 atm (9,500 kPa)
(1,938.00 K)

Chlorine 143.8 °C (416.9 K) 76.0 atm (7,700 kPa)

Ethanol (C2H5OH) 241 °C (514 K) 62.18 atm (6,300 kPa)

Fluorine −128.85 °C (144.30 K) 51.5 atm (5,220 kPa)

Helium −267.96 °C (5.19 K) 2.24 atm (227 kPa)

Hydrogen −239.95 °C (33.20 K) 12.8 atm (1,300 kPa)

Krypton −63.8 °C (209.3 K) 54.3 atm (5,500 kPa)

Methane (CH4) −82.3 °C (190.8 K) 45.79 atm (4,640 kPa)

Neon −228.75 °C (44.40 K) 27.2 atm (2,760 kPa)

Nitrogen −146.9 °C (126.2 K) 33.5 atm (3,390 kPa)


Substance[15][16] Critical temperature Critical pressure (absolute)

Oxygen −118.6 °C (154.6 K) 49.8 atm (5,050 kPa)

Carbon dioxide (CO2) 31.04 °C (304.19 K) 72.8 atm (7,380 kPa)

Nitrous oxide (N2O) 36.4 °C (309.5 K) 71.5 atm (7,240 kPa)

Sulfuric acid (H2SO4) 654 °C (927 K) 45.4 atm (4,600 kPa)

Xenon 16.6 °C (289.8 K) 57.6 atm (5,840 kPa)

Lithium 2,950 °C (3,220 K) 652 atm (66,100 kPa)

Mercury 1,476.9 °C (1,750.1 K) 1,720 atm (174,000 kPa)

1,040.85 °C
Sulfur 207 atm (21,000 kPa)
(1,314.00 K)

Iron 8,227 °C (8,500 K)

Gold 6,977 °C (7,250 K) 5,000 atm (510,000 kPa)

Aluminium 7,577 °C (7,850 K)

Water (H2O)[2][18] 373.946 °C (647.096 K) 217.7 atm (22,060 kPa)

Mixtures: liquid–liquid critical point [edit]


A plot of typical polymer solution phase behavior including two critical points: an LCST and a UCST.

The liquid–liquid critical point of a solution, which occurs at the critical solution temperature, occurs
at the limit of the two-phase region of the phase diagram. In other words, it is the point at which an
infinitesimal change in some thermodynamic variable (such as temperature or pressure) will lead to
separation of the mixture into two distinct liquid phases, as shown in the polymer–solvent phase
diagram to the right. Two types of liquid–liquid critical points are the upper critical solution
temperature (UCST), which is the hottest point at which cooling will induce phase separation, and
the lower critical solution temperature (LCST), which is the coldest point at which heating will
induce phase separation.

Mathematical definition[edit]
From a theoretical standpoint, the liquid–liquid critical point represents the temperature-
concentration extremum of the spinodal curve (as can be seen in the figure to the right). Thus, the
liquid–liquid critical point in a two-component system must satisfy two conditions: the condition of
the spinodal curve (the second derivative of the free energy with respect to concentration must
equal zero), and the extremum condition (the third derivative of the free energy with respect to
concentration must also equal zero or the derivative of the spinodal temperature with respect to
concentration must equal zero).
Real Gases

At low temperatures and high pressures, the ideal gas equation ceases to apply. Under these conditions, you must deal with "real" gases.

A quick way of checking the validity of an ideal gas assumption is to look at the specific molar volume of the gas. The ideal gas equation is good to
within about 1% if:

Real gases require more complex equations of state than do ideal gases. There are many options, but we will focus on:

 Virial Equations of State


 Cubic Equations of State
 Compressibility Factor Equation

This topic will be an important part of Thermo II.

Virial Equations of State

Virial equations are a family of equations of state of the general form:

The parameters in the equation (B,C,D = ci) are called "virial coefficients". If ci=0 for i>0, the virial equation reduces to the ideal gas equation. Just as
with the ideal gas equation, the temperatures and pressures used must be absolute.

The accuracy required determines the number of terms that are kept -- more terms makes the equation more accurate, but also more complicated to
work with. Virial coefficients are different for each gas, but other than that are functions of temperature only.

Coefficients are normally obtained by making measurements of P, V, and T, and fitting the equation. These values are then published so that others
may use them.

Many forms of the virial equation exist. Often, we will truncate the virial equation to

A number of methods (correlations, etc.) are available to determine B. In order to improve accuracy and capture more behaviors, additional
parameters are sometimes added. One example is the Benedict-Webb-Rubin (BWR) equation of state.

All the constants must be supplied if you are to use this equation for a particular gas. It isn't always easy to find BWR coefficients for the gas you are
interested in.
Cubic Equations of State

Virial equations cannot represent thermodynamic systems where both liquid and vapor are present. A "cubic" EoS is need to do this. One such is the
Soave-Redlich-Kwong (SRK) equation.

where the constants are given by

In this equation, the b term is a volume correction, while the a is a molecular interaction parameter. The constants all depend on the critical
temperature and pressure of the gas. These can be looked up easily in a data table.

The "acentric factor", omega, is also easily looked up. It is related to the geometry of the gas molecule.

To use the SRK equation:

1. look up Tc, Pc, and the acentric factor


2. plug in and find a, b, and alpha
3. plug these into the SRK equation; the result will be a cubic equation in P, T, and V
4. solve for the unknown you seek

Solving the cubic equation typically requires an iterative ("trial-and- error") solution; most of the time you probably want to use a computer or
calculator routine for this task.

EXAMPLE:
Carbon dioxide at 300 K and 6.8 atm flows at 100 kmol/h. Use the SRK equation of state to determine the volumetric flow. (Felder & Rousseau,
1986, Example 5.3-3)

Strategy: The pressure and temperature are known, so look up the critical properties and acentric factor, find the SRK equation constants, and solve
the SRK equation for the specific molar volume.

The critical properties (from the back of the book) are

so the constants become


Inserting these into the SRK equation gives

All of the terms have units of atm. The equation can be rearranged to make solution more convenient

Solution of the Equation: This equation is cubic in V, so trying to solve it algebraically is a bit complicated. Consequently, we will use an iterative
approach.

We can implement the iterative strategy in a number of ways -- and using a computer or a calculator solver is strongly recommended. All the
approaches use the same basic approach. A value of V is guessed and the right hand side (RHS) of the equation calculated. The result will be
compared to zero and a new value of V chosen and tried.

The ideal gas law is always a good place to get an initial value, because it will usually be "close" to the real solution:

You also need to determine how close to the "exact" answer you want to be. This error tolerance is usually built-in to the solver routine, although you
usually can adjust it (in Mathcad, look at the value of TOL). If you've got a particularly difficult system, sometimes it is useful to relax the tolerance
to get a rough solution, then update your estimate, tighten the tolerance, and repeat the solution for a final value.

This is the type of problem that Mathcad handles very well:


Critical Point

This module refers to a finite amount of particles placed in a closed container (i.e. no volume change) in which boiling
cannot occur. The inability for boiling to occur- because the particles in the container are not exposed to the atmosphere,
results in the incessant increase of temperature and pressure. The critical point is the temperature and pressure at which
the distinction between liquid and gas can no longer be made.

Introduction

At the critical point, the particles in a closed container are thought to be vaporizing at such a rapid rate that the density of
liquid and vapor are equal, and thus form a supercritical fluid. As a result of the high rates of change, the surface tension
of the liquid eventually disappears. You will have noticed that this liquid-vapor equilibrium curve has a top limit (labeled
as C in the phase diagram in Figure 1), which is known as the critical point. The temperature and pressure corresponding
to this are known as the critical temperature and critical pressure. If you increase the pressure on a gas (vapor) at a
temperature lower than the critical temperature, you will eventually cross the liquid-vapor equilibrium line and the vapor
will condense to give a liquid.

Figure 1: Phase diagram for a single component system  with critical point emphasized

This works fine as long as the gas is below the critical temperature. What, though, if your
temperature was above the critical temperature? There wouldn't be any line to cross! That is
because, above the critical temperature, it is impossible to condense a gas into a liquid just by
increasing the pressure. All you get is a highly compressed gas. The particles have too much
energy for the intermolecular attractions to hold them together as a liquid. The critical
temperature obviously varies from substance to substance and depends on the strength of the
attractions between the particles. The stronger the intermolecular attractions, the higher the
critical temperature.

Why the Critical Point is Important

The condensation of a gas will never occur above the critical point. A massive amount of pressure can be applied to a gas
in a closed container, and it may become highly dense, but will not exhibit a meniscus. Molecules at critical temperatures
possess high kinetic energy, and as a result the intermolecular forces in the molecules are weakened.

The Declined Critical Points of Polymer Solutions

A novel discovery made by the University of Manchester, identified that lower critical temperatures are existent in polymer
solutions. It has been manifested that hydrocarbon polymers integrated with a hydrocarbon solution portrays what the
university terms a "L.C.S.T." or a lower critical solution temperature. This lower critical solution temperature of polymers
has been proclaimed to be in a range near the gas- liquid critical point of the polymer's solvent, and can reach up to 170
degrees Celsius. Such a lower critical solution temperature can be contributed to the assimilation of the heat and volume
of the substance n-pentane with most hydrocarbon polymers at room temperature (Freeman, P.I., Rowlinson, J.S.).

The Effects of Wetting on the Critical Point

When a fluid is present in two phases, in a container, and a critical point is near establishment, contact with the
imminently forming third phase does not occur. This phenomena can be accounted for by examining the other two
existing phases; the third phase does not immediately form because one of the other two phases wets the third phase,
causing it to be eliminated. This wetting phase will continually occur when a phase is not entirely stable as a whole.

Problems

Temperature and vapor pressure are essential to the stimulation of a critical point; the following problems interconnect
the two concepts.

1. Determine the vapor pressure of liquid gold at 1936 °C if 200 L of Ne gas is permitted to assimilate with the
metal causing it to decline in mass by 0.213 g.
2. Approximately 0.423 g H2O is preserved at a temperature of 62 °C in a sealed flask at a volume of 0.726 L. When
in equilibrium, will the water be present solely as a liquid?
3. Taking into account the previous problem, would it be probable for the sample of water in the flask to be solitarily
present as a vapor?
4. Is it possible for the the sample of water in problem number two to exist in both the liquid and vapor form at
equilibrium?
5. The vapor pressure of H2O at 30 °C is 46.2 mmHg. What is the vapor pressure of water at 28 °C, if the enthalpy
of fusion is 44 kJ/mol?

Solutions

1. In correlation with Dalton's Law of partial pressure, the gold vapor can be interpreted as a solitary gas occurring at the
given volume of Neon. The ideal-gas law will be utilized to determine the vapor pressure.

PV=nRT(1)(1)PV=nRT

 P=?
 V=200L
 T=1936.C (to convert to K add 273)=2209 K
 R=00820L atm/mol K

Convert to mol the molar mass of Au

n=0.213g(1molAu196.9gAu)=0.00108molAu(2)(2)n=0.213g(1molAu196.9gAu)=0.00108molAu

PV=nRTPV=nRT can then be algebraically converted into P=nRTVP=nRTV. Substitute you attained values.

P=(0.00108mol)×(0.08206Latm/molK)×(2209K)200L(3)(3)P=(0.00108mol)×(0.08206Latm/molK)×(2209K)200L

P=9.78×10−4atm(4)(4)P=9.78×10−4atm

2. Taking into account the density of water, 1 g/ml, a 0.423 g sample of H2OH2O would amount to 0.42 ml. Considering
this, it can be concluded that the water in the flask could not solely exist as a liquid at equilibrium since a 0.42 ml sample
of water is incapable of accounting for the overall volume present in the flask (0.726L).
3. The ideal gas law can be applied to determine whether or not the water in the flask could solely exist in the vapor
form. Rearrange the Ideal Gas Law:

P=nRTV(5)(5)P=nRTV

 n= (0.423g)*(1 mol/18 g H2O)=0.0235mol


 R=0.08206 L atm/ mol K
 T=62°C + 273 = 335K
 V=0.726 L
P=(0.0235mol)×(0.08206Latm/molK)×(335K)0.726L(6)(6)P=(0.0235mol)×(0.08206Latm/molK)×(335K)0.726L

P=0.889atm(7)(7)P=0.889atm

The actual Vapor pressure of water at 60 °C is 149.4mmHg149.4mmHg.

To make a comparison to the vapor pressure attained at 62°C (which is within a close range of the temperature of 60
degrees Celsius), convert the vapor pressure calculated in atm to mmHg.

(0.889atm)∗(760mmHg1atm)=676.27mmHg(8)(8)(0.889atm)∗(760mmHg1atm)=676.27mmHg

Comparing this vapor pressure with the actual vapor pressure of water at 60 °C, it can be concluded that it is improbable
that water solely exists in the flask as vapor since this attained vapor pressure exceeds that of the actual vapor pressure
that occurs naturally.

4. As the prior explications and calculations have proven, the sample of water is incapable of existing in the flask as either
a liquid or vapor alone. Therefore vapor and liquid must exist together at 60 °C and 149.4 mmHg.

5. To evaluate the pressure of water at 28 °C, the Clausius-Clapeyron equation must be used.

lnP1P2=HvapR(1T2−1T1)(9)(9)ln⁡P1P2=HvapR(1T2−1T1)

Begin by designating the variables in the equation to the given values in the problem.

If T2=30°CT2=30°C and P2=46.2mmHgP2=46.2mmHg, what is P1P1 when T1=28°CT1=28°C

The units for temperature must be in Kelvin.

 T2=30°C+273.15K=303.15KT2=30°C+273.15K=303.15K
 T1=28°C+273.15=301.15KT1=28°C+273.15=301.15K

The enthalpy of vaporization (44 kJ/ mol) should be converted to J/ mol to allow for unit cancelation.

44kJ/mol(1000J1kJ)=44.0×103J/mol.(10)(10)44kJ/mol(1000J1kJ)=44.0×103J/mol.

R= 8.3145 J/mol K

Substituting the values in the equation one you obtain:

ln(46.2mmHgP1)=44,000Jmol−18.3145J/molK(1301.15K−1303.15K)(11)
(11)ln⁡(46.2mmHgP1)=44,000Jmol−18.3145J/molK(1301.15K−1303.15K)
To eliminate the natural logarithm, take the exponentl of both sides:

eln⎛⎝46.2mmHgP1⎞⎠=e44,000Jmol−18.3145J/molK(1301.15K−1303.15K)(12)

(12)eln⁡(46.2mmHgP1)=e44,000Jmol−18.3145J/molK(1301.15K−1303.15K)

46.2mmHgP1=e44,000Jmol−18.3145J/molK(1301.15K−1303.15K)(13)

(13)46.2mmHgP1=e44,000Jmol−18.3145J/molK(1301.15K−1303.15K)

46.2mmHgP1=e0.11593(14)(14)46.2mmHgP1=e0.11593

46.2mmHgP1=1.1229(15)(15)46.2mmHgP1=1.1229

1P1=1.122946.2mmHg(16)(16)1P1=1.122946.2mmHg

P1=46.2mmHg1.1229(17)(17)P1=46.2mmHg1.1229

P1=41.143mmHgP1=41.143mmHg, which gets rounded to two significant digits (from the two significant digits in 44
kJ/mol).

P1P1 vapor pressure is 41 mmHg.


Amagat's law
From Wikipedia, the free encyclopedia

Jump to navigationJump to search

Amagat's law or the Law of Partial Volumes describes the behaviour and properties of mixtures of ideal (as well as
some cases of non-ideal) gases. Of use in chemistry and thermodynamics.

Overview[edit]
Amagat's law states that the extensive volume V = N·v of a gas mixture is equal to the sum of volumes Vi of
the K component gases, if the temperature T and the pressure p remain the same:[1][2]
This is the experimental expression of volume as an extensive quantity. It is named after Emile Amagat.
According to Amagat's law of partial volume, the total volume of a non-reacting mixture of gases at constant
temperature and pressure should be equal to the sum of the individual partial volumes of the constituent gases. So
if  are considered to be the partial volumes of components in the gaseous mixture, then the total volume  would be
represented as:
Both Amagat's and Dalton's laws predict the properties of gas mixtures. Their predictions are the same for ideal
gases. However, for real (non-ideal) gases, the results differ. [3] Dalton's law of partial pressures assumes that the
gases in the mixture are non-interacting (with each other) and each gas independently applies its own pressure,
the sum of which is the total pressure. Amagat's law assumes that the volumes of the component gases (again
at the same temperature and pressure) are additive; the interactions of the different gases are the same as the
average interactions of the components.
The interactions can be interpreted in terms of a second virial coefficient, B(T), for the mixture. For two
components, the second virial coefficient for the mixture can be expressed as:
where the subscripts refer to components 1 and 2, the X's are the mole fractions, and the B's are the
second virial coefficients. The cross term, B1,2, of the mixture is given by:
 (Dalton's law)
and
 (Amagat's law).
When the volumes of each component gas (same temperature and pressure) are very similar, then
Amagat's law becomes mathematically equivalent to Vegard's law for solid mixtures.

Ideal gas mixture[edit]


When Amagat's law is valid and the gas mixture is made of ideal gases (Ideal gas law):
where:
  is the pressure of the gas mixture,
  is the volume of the i-component of the gas mixture,
  is the total volume of the gas mixture,
  is the amount of substance of i-component of the gas mixture (in mol),
  is the total amount of substance of gas mixture (in mol),
  is the ideal, or universal, gas constant, equal to the product of the Boltzmann constant and
the Avogadro constant,
  is the absolute temperature of the gas mixture (in K),
  is the mole fraction of the i-component of the gas mixture.
Dalton's law
From Wikipedia, the free encyclopedia

Jump to navigationJump to search

For the law of stoichiometry, see  Law of multiple proportions.

An illustration of Dalton's law using the gases of air at sea level.

In chemistry and physics, Dalton's law (also called Dalton's law of partial pressures) states that in a mixture of non-
reacting gases, the total pressure exerted is equal to the sum of the partial pressures of the individual gases.
[1]
 This empirical law was observed by John Dalton in 1801 and published in 1802.[2] Dalton's law is related to the ideal gas
laws.

Contents

 1Formula
 2Volume-based concentration
 3See also
 4References

Formula[edit]
Mathematically, the pressure of a mixture of non-reactive gases can be defined as the summation:
       or      
where p1, p2, ..., pn represent the partial pressures of each component. [1]
where xi is the mole fraction of the ith component in the total mixture of n components.

Volume-based concentration[edit]
The relationship below provides a way to determine the volume-based concentration of any individual gaseous
component
where ci is the concentration of the ith component.
Dalton's law is not strictly followed by real gases, with the deviation increasing with pressure. Under such
conditions the volume occupied by the molecules becomes significant compared to the free space between
them. In particular, the short average distances between molecules increases intermolecular forces between
gas molecules enough to substantially change the pressure exerted by them, an effect not included in the
ideal gas model.

Das könnte Ihnen auch gefallen