Sie sind auf Seite 1von 36

Van 't Hoff equation

The Van't Hoff equation relates the change


in the equilibrium constant, Keq, of a
chemical reaction to the change in
temperature, T, given the standard
enthalpy change, ΔH⊖, for the process. It
was proposed by Dutch chemist Jacobus
Henricus van 't Hoff in 1884 in his book
"Études de Dynamique chimique" (Studies
in dynamic chemistry).[1] This equation is
sometimes also referred to as the
Vukancic–Vukovic equation.[2][3][4]

The Van 't Hoff equation has been widely


utilized to explore the changes in state
functions in a thermodynamic system. The
Van 't Hoff plot, which is derived from this
equation, is especially effective in
estimating the change in enthalpy, or total
energy, and entropy, or amount of disorder,
of a chemical reaction

Equation
Under standard conditions
Under standard conditions, the Van 't Hoff
equation is[5][6]

where R is the ideal gas constant. This


equation is exact at any one temperature.
In practice, the equation is often integrated
between two temperatures under the
assumption that the reaction enthalpy ΔH
is constant. Since in reality ΔH and ΔS do
vary with temperature for most
processes,[7] the integrated equation is
only approximate.
A major use of the integrated equation is
to estimate a new equilibrium constant at
a new absolute temperature assuming a
constant standard enthalpy change over
the temperature range.

To obtain the integrated equation, it is


convenient to first rewrite the Van 't Hoff
equation as[5]

The definite integral between


temperatures T1 and T2 is then
In this equation K1 is the equilibrium
constant at absolute temperature T1, and
K2 is the equilibrium constant at absolute
temperature T2.

Development from
thermodynamics

Combining the definition of Gibbs free


energy:

where S is the entropy of the system, and


the Gibbs free energy isotherm equation:[8]
we obtain

Differentiation of this expression with


respect to the variable 1/T yields the Van 't
Hoff equation.

Provided that ΔH⊖ and ΔS⊖ are constant,


the preceding equation gives ln K as a
linear function of T1 and hence is known as
the linear form of the Van 't Hoff equation.
Therefore, when the range in temperature
is small enough that the standard enthalpy
and entropy changes are essentially
constant, a plot of the natural logarithm of
the equilibrium constant versus the
reciprocal temperature gives a straight
line. The slope of the line may be
multiplied by the gas constant R to obtain
the standard enthalpy change of the
reaction, and the intercept may be
multiplied by R to obtain the standard
entropy change.

Van 't Hoff isotherm

The Gibbs free energy changes with the


temperature and pressure of the
thermodynamic system. The Van 't Hoff
isotherm can be used to determine the
Gibbs free energy for non-standard state
reactions at a constant temperature:[9]

where ΔrG is the Gibbs free energy for the


reaction, and Qr is the reaction quotient.
When a reaction is at equilibrium, Qr = Keq.
The Van 't Hoff isotherm can help estimate
the equilibrium reaction shift. When
ΔrG < 0, the reaction moves in the forward
direction. When ΔrG > 0, the reaction
moves in the backwards directions. See
Chemical equilibrium.

Van 't Hoff plot


For a reversible reaction, the equilibrium
constant can be measured at a variety of
temperatures. This data can be plotted on
a graph with ln Keq on the y axis and T1 on
the x axis. The data should have a linear
relationship, the equation for which can be
found by fitting the data using the linear
form of the Van 't Hoff equation

This graph is called the Van 't Hoff plot


and is widely used to estimate the
enthalpy and entropy of a chemical
reaction. From this plot, − ΔH
R
is the slope,
and ΔS
R is the intercept of the linear fit.
By measuring the equilibrium constant,
Keq, at different temperatures, the Van 't
Hoff plot can be used to assess a reaction
when temperature changes.[10][11] Knowing
the slope and intercept from the Van 't
Hoff plot, the enthalpy and entropy of a
reaction can be easily obtained using

The Van 't Hoff plot can be used to quickly


determine the enthalpy of a chemical
reaction both qualitatively and
quantitatively. Change in enthalpy can be
positive or negative, leading to two major
forms of the Van 't Hoff plot.
Endothermic reactions

Van 't Hoff plot for an endothermic reaction

For an endothermic reaction, heat is


absorbed, making the net enthalpy change
positive. Thus, according to the definition
of the slope:
for an endothermic reaction, ΔH > 0 (the
gas constant R > 0), so

Thus, for an endothermic reaction, the Van


't Hoff plot should always have a negative
slope.

Exothermic reactions
Van 't Hoff plot for an exothermic reaction

For an exothermic reaction, heat is


released, making the net enthalpy change
negative. Thus, according to the definition
of the slope:

fom an exothermic reaction, ΔH < 0, so


Thus, for an exothermic reaction, the Van 't
Hoff plot should always have a positive
slope.

Error propagation

Using the fact that ΔG⊖ = −RT ln K =


ΔH⊖ − TΔS⊖, it would appear that two
measurements of K would suffice to be
able to obtain a value of ΔH⊖:
where K1 and K2 are the equilibrium
constant values obtained at temperatures
T1 and T2 respectively. The precision of
ΔH⊖ values obtained in this way is highly
dependent on the precision of the
equilibrium constant values. A typical pair
of temperatures might be 25 and 35 °C.
For these temperatures

Inserting this value into the expression for


ΔH⊖:
Now, error propagation shows that the
error on ΔH⊖ will be about 76 kJ/mol
times the error on (ln K1 − ln K2), or about
110 kJ/mol times the error on the ln K
values. Assume for example that the error
on each ln K, is σ ≈ 0.05, a small but
reasonable value. The error on ΔH⊖ will be
about 5 kJ/mol. So, even though the
individual stability constants were
determined with good precision, the
enthalpy calculated in this way is subject
to a significant error.

The entropy will then be obtained from


ΔS⊖ =  T1 (ΔH⊖ + RT ln K). In this expression
the error on the second term is negligible
compared to the error on the first term.
The magnifying factor is then
76 kJ/mol ÷ 298 K, so for an error of 0.05
in the logarithms the error on ΔS⊖ will be
of the order of 17 J/(K·mol).

When equilibrium constants are measured


at three or more temperatures, values of
ΔH⊖ will be obtained by straight-line
fitting. In this case the error on the
standard enthalpy will be magnified to a
somewhat lesser, but still substantial,
extent.

Applications of the Van 't Hoff


plot
Van 't Hoff analysis

Van 't Hoff analysis

In biological research, the Van 't Hoff plot


is also called Van 't Hoff analysis.[12] It is
most effective in determining the favored
product in a reaction.
Assume two products B and C form in a
reaction:

a A + d D → b B,
a A + d D → c C.

In this case, Keq can be defined as ratio of


B to C rather than the equilibrium
constant.

B
When C  > 1, B is the favored product, and
the data on the Van 't Hoff plot will be in
the positive region.

B  < 1, C is the favored product, and


When C
the data on the Van 't Hoff plot will be in
the negative region.
Using this information, a Van 't Hoff
analysis can help determine the most
suitable temperature for a favored
product.

Recently, a Van 't Hoff analysis was used


to determine whether water preferentially
forms a hydrogen bond with the C-
terminus or the N-terminus of the amino
acid proline.[13] The equilibrium constant
for each reaction was found at a variety of
temperatures, and a Van 't Hoff plot was
created. This analysis showed that
enthalpically, the water preferred to
hydrogen bond to the C-terminus, but
entropically it was more favorable to
hydrogen bond with the N-terminus.
Specifically, they found that C-terminus
hydrogen bonding was favored by 4.2–
6.4 kJ/mol. The N-terminus hydrogen
bonding was favored by 31–43 J/(K·mol).

This data alone could not conclude which


site water will preferentially hydrogen-bond
to, so additional experiments were used. It
was determined that at lower
temperatures, the enthalpically favored
species, the water hydrogen-bonded to the
C-terminus, was preferred. At higher
temperatures, the entropically favored
species, the water hydrogen-bonded to the
N-terminus, was preferred.
Mechanistic studies

Van 't Hoff plot in mechanism study

A chemical reaction may undergo different


reaction mechanisms at different
temperatures.[14]

In this case, a Van 't Hoff plot with two or


more linear fits may be exploited. Each
linear fit has a different slope and
intercept, which indicates different
changes in enthalpy and entropy for each
distinct mechanisms. The Van 't Hoff plot
can be used to find the enthalpy and
entropy change for each mechanism and
the favored mechanism under different
temperatures.

In the example figure, the reaction


undergoes mechanism 1 at high
temperature and mechanism 2 at low
temperature.

Temperature dependence

Temperature-dependent Van 't Hoff plot

The Van 't Hoff plot is linear based on the


assumption that the enthalpy and entropy
are constant with temperature changes.
However, in some cases the enthalpy and
entropy do change dramatically with
temperature. A first-order approximation is
to assume that the two different reaction
products have different heat capacities.
Incorporating this assumption yields an
c
additional term T2 in the expression for the
equilibrium constant as a function of
temperature. A polynomial fit can then be
used to analyze data that exhibits a non-
constant standard enthalpy of reaction:[15]

where
Thus, the enthalpy and entropy of a
reaction can still be determined at specific
temperatures even when a temperature
dependence exists.

Surfactant self-assembly

The Van 't Hoff relation is particularly


useful for the determination of the

micellization enthalpy ΔHm of surfactants
from the temperature dependence of the
critical micelle concentration (CMC):
However, the relation loses its validity
when the aggregation number is also
temperature-dependent, and the following
relation should be used instead:[16]

with GN+1 and GN being the free energies


of the surfactant in a micelle with
aggregation number N + 1 and N
respectively. This effect is particularly
relevant for nonionic ethoxylated
surfactants[17] or polyoxypropylene–
polyoxyethylene block copolymers
(Poloxamers, Pluronics, Synperonics).[18]
The extended equation can be exploited
for the extraction of aggregation numbers
of self-assembled micelles from
differential scanning calorimetric
thermograms.[19]

See also
Clausius–Clapeyron relation
Van 't Hoff factor (i)
Gibbs–Helmholtz equation

References
1. Biography on Nobel prize website .
Nobelprize.org (1911-03-01). Retrieved on
2013-11-08.
2. Journal on Modeling Prograde TiO2
Activity and its significance for Ti in Quartz
Thermobarometry of Pelitic Metamorphic
Rocks . Academia.edu. p. 2.
3. Journal on Indirect spectrophotometric
determination of folic acid based on the
oxidation reaction and studying some of the
thermodynamic parameters .
Academia.edu. p. 67.
4. PNAS Supporting Information
Correction . (PNAS) Proceedings of the
National Academy of Sciences of the United
States of America. November 30, 2012. p. 3.
5. Atkins, Peter; De Paula, Julio (10 March
2006). Physical Chemistry (8th ed.). W. H.
Freeman and Company. p. 212. ISBN 0-
7167-8759-8.
6. Ives, D. J. G. (1971). Chemical
Thermodynamics. University Chemistry.
Macdonald Technical and Scientific. ISBN 0-
356-03736-3.
7. Craig, Norman (1996). "Entropy
Diagrams" . J. Chem. Educ. 73: 710.
Bibcode:1996JChEd..73..710C .
doi:10.1021/ed073p710 .
8. Dickerson, R. E.; Geis, I. (1976).
Chemistry, Matter, and the Universe. USA:
W. A. Benjamin Inc. ISBN 0-19-855148-7.
9. Monk, Paul (2004). Physical Chemistry:
Understanding our Chemical World. Wiley.
p. 162. ISBN 978-0471491811.
10. Kim, Tae Woo (2012). "Dynamic
[2]Catenation of Pd(II) Self-assembled
Macrocycles in Water". Chem. Lett. 41: 70.
doi:10.1246/cl.2012.70 .
11. Ichikawa, Takayuki (2010).
"Thermodynamic properties of metal
amides determined by ammonia pressure-
composition isotherms". J. Chem.
Thermodynamics. 42: 140.
doi:10.1016/j.jct.2009.07.024 .
12. "Van't Hoff Analysis" . Protein Analysis
and Design Group.
13. Prell, James; Williams E. (2010).
"Entropy Drives an Attached Water Molecule
from the C- to N-Terminus on Protonated
Proline". J. Am. Chem. Soc. 132 (42):
14733. doi:10.1021/ja106167d .
PMID 20886878 .
14. Chatake, Toshiyuki (2010). "An
Approach to DNA Crystallization Using the
Thermal Reversible Process of DNA
Duplexes". Cryst. Growth Des. 10: 1090.
doi:10.1021/cg9007075 .
15. David, Victor (28 April 2011). "Deviation
from van't Hoff dependence in RP-LC
induced by tautomeric interconversion
observed for four compounds". Journal of
Separation Science. 34 (12): 1423.
doi:10.1002/jssc.201100029 .
16. Holtzer, Alfred; Holtzer, Marilyn F. (2002-
05-01). "Use of the van't Hoff relation in
determination of the enthalpy of micelle
formation" . The Journal of Physical
Chemistry. 78 (14): 1442–1443.
doi:10.1021/j100607a026 .
17. Heerklotz, Heiko; Tsamaloukas, Alekos;
Kita-Tokarczyk, Katarzyna; Strunz, Pavel;
Gutberlet, Thomas (2004-11-25). "Structural,
Volumetric, and Thermodynamic
Characterization of a Micellar Sphere-to-
Rod Transition" . Journal of the American
Chemical Society. 126 (50): 16544–16552.
doi:10.1021/ja045525w . PMID 15600359 .
18. Taboada, Pablo; Mosquera, Victor;
Attwood, David; Yang, Zhuo; Booth, Colin.
"Enthalpy of micellisation of a diblock
copoly(oxyethylene/oxypropylene) by
isothermal titration calorimetry. Comparison
with the van't Hoff value" . Physical
Chemistry Chemical Physics. 5 (12): 2625.
Bibcode:2003PCCP....5.2625T .
doi:10.1039/b303108j .
19. Chiappisi, Leonardo; Lazzara, Giuseppe;
Gradzielski, Michael; Milioto, Stefana (2012-
12-06). "Quantitative Description of
Temperature Induced Self-Aggregation
Thermograms Determined by Differential
Scanning Calorimetry" . Langmuir. 28 (51):
17609–17616. doi:10.1021/la303599d .
Retrieved from
"https://en.wikipedia.org/w/index.php?
title=Van_%27t_Hoff_equation&oldid=819090709"

Last edited 24 days ago by an anon…

Content is available under CC BY-SA 3.0 unless


otherwise noted.

Das könnte Ihnen auch gefallen