Sei sulla pagina 1di 18

Enthalpy of mixing

The enthalpy of mixing (also called heat of mixing) is the heat that is taken up or released upon
mixing of two (non-reacting) chemical substances. When the enthalpy of mixing is positive,
mixing is endothermic while negative enthalpy of mixing signifies exothermic mixing. In ideal
mixtures the enthalpy of mixing is null. In non-ideal mixtures the thermodynamic activity of
each component is different from its concentration by multiplying with the activity coefficient.
The enthalpy of solution, enthalpy of dissolution, or heat of solution is the enthalpy change
associated with the dissolution of a substance in a solvent at constant pressure resulting in
infinite dilution.
The enthalpy of solution is most often expressed in kJ/mol at constant temperature. The energy
change can be regarded as being made of three parts, the endothermic breaking of bonds within
the solute and within the solvent, and the formation of attractions between the solute and the
solvent. An ideal solution has an enthalpy of solution of zero. For a non-ideal solution it is an
excess molar quantity.
Dissolution by most gases is exothermic. That is, when a gas dissolves in a liquid
solvent, energy is released as heat, warming both the system (i.e. the solution) and
the surroundings. The temperature of the solution then decreases to that of the
surroundings. The equilibrium, between the gas as a separate phase and the gas in
solution, will therefore (by Le Chtelier's principle) shift to favour the gas going into
solution as the temperature is decreased. Thus, decreasing the temperature
increases the solubility of a gas. When a saturated solution of a gas is heated, gas
comes out of solution.

Steps in dissolution
Dissolution can be viewed as occurring in three steps:
1. Breaking solute-solute attractions (endothermic), see for instance lattice
energy in salts.
2. Breaking solvent-solvent attractions (endothermic), for instance that of
hydrogen bonding
3. Forming solvent-solute attractions (exothermic), in solvation.

The value of the enthalpy of solution is the sum of these individual steps. Dissolving ammonium
nitrate in water is endothermic. The energy released by solvation of the ammonium ions and
nitrate ions is less than the energy absorbed in breaking up the ammonium nitrate ionic lattice

and the attractions between water molecules. Dissolving potassium hydroxide is exothermic, as
more energy is released during solvation than is used in breaking up the solute and solvent.

Dependence on the nature of the solution


The enthalpy of solution of an ideal solution is zero by definition. For non-ideal solutions it is
connected to the activity coefficient of the solute(s) and the temperature derivative of the relative
permittivity.

Enthalpy change of solution for some selected compounds


hydrochloric acid

-74.84

ammonium nitrate

+25.69

ammonia

-30.50

potassium hydroxide

-57.61

caesium hydroxide

-71.55

sodium chloride

+3.87

potassium chlorate

+41.38

acetic acid

-1.51

sodium hydroxide

-44.51

Change in enthalpy Ho in kJ/mol in water at 25C

Ideal solution

In chemistry, an ideal solution or ideal mixture is a solution with thermodynamic properties


analogous to those of a mixture of ideal gases. The enthalpy of solution (or "enthalpy of mixing")
is zero as is the volume change on mixing; the closer to zero the enthalpy of solution is, the more
"ideal" the behavior of the solution becomes. The vapor pressure of the solution obeys Raoult's
law, and the activity coefficient of each component (which measures deviation from ideality) is
equal to one.
The concept of an ideal solution is fundamental to chemical thermodynamics and its
applications, such as the use of colligative properties.

Physical origin
Ideality of solutions is analogous to ideality for gases, with the important difference that
intermolecular interactions in liquids are strong and cannot simply be neglected as they can for
ideal gases. Instead we assume that the mean strength of the interactions are the same between
all the molecules of the solution.
More formally, for a mix of molecules of A and B, the interactions between unlike neighbors
(UAB) and like neighbors UAA and UBB must be of the same average strength i.e. 2 UAB = UAA +
UBB and the longer-range interactions must be nil (or at least indistinguishable). If the molecular
forces are the same between AA, AB and BB, i.e. UAB = UAA = UBB, then the solution is
automatically ideal.
If the molecules are almost identical chemically, e.g. 1-butanol and 2-butanol, then the solution
will be almost ideal. Since the interaction energies between A and B are almost equal, it follows
that there is a very small overall energy (enthalpy) change when the substances are mixed. The
more dissimilar the nature of A and B, the more strongly the solution is expected to deviate from
ideality.

Formal definition
Different related definitions of an ideal solution have been proposed. The simplest definition is
that an ideal solution is a solution for which each component (i) obeys Raoult's law
for all compositions. Here is the vapor pressure of component i above the solution,
mole fraction and

is its

is the vapor pressure of the pure substance i at the same temperature.[3][4][5]

This definition depends on vapor pressures which are a directly measurable property, at least for
volatile components. The thermodynamic properties may then be obtained from the chemical
potential (or partial molar Gibbs energy g) of each component, which is assumed to be given
by the ideal gas formula

The reference pressure


operations.

may be taken as

On substituting the value of

= 1 bar, or as the pressure of the mix to ease

from Raoult's law,

This equation for the chemical potential can be used as an alternate definition for an ideal
solution.
However, the vapor above the solution may not actually behave as a mixture of ideal gases.
Some authors therefore define an ideal solution as one for which each component obeys the
fugacity analogue of Raoult's law

,
is the fugacity of as a pure substance.[6]

Here is the fugacity of component in solution and


[7]
Since the fugacity is defined by the equation

this definition leads to ideal values of the chemical potential and other thermodynamic properties
even when the component vapors above the solution are not ideal gases. An equivalent statement
uses thermodynamic activity instead of fugacity.[8]

Thermodynamic properties
Volume

If we differentiate this last equation with respect to

at

constant we get:

but we know from the Gibbs potential equation that:

These last two equations put together give:

Since all this, done as a pure substance is valid in a mix just adding the subscript to all the
intensive variables and changing to , standing for Partial molar volume.

Applying the first equation of this section to this last equation we get

which means that in an ideal mix the volume is the addition of the volumes of its components.
Enthalpy and heat capacity

Proceeding in a similar way but derivative with respect of


enthalpies

derivative with respect to T and remembering that

which in turn is

we get to a similar result with

we get:

Meaning that the enthalpy of the mix is equal to the sum of its components.
Since

and

It is also easily verifiable that

Entropy of mixing

Finally since

Which means that

and since

then

At last we can calculate the entropy of mixing since

and

Consequences
Solvent-Solute interactions are similar to solute-solute and solvent-solvent interactions
Since the enthalpy of mixing (solution) is zero, the change in Gibbs free energy on mixing is
determined solely by the entropy of mixing. Hence the molar Gibbs free energy of mixing is

or for a two component solution

where m denotes molar i.e. change in Gibbs free energy per mole of solution, and
fraction of component .

is the mole

Note that this free energy of mixing is always negative (since each is positive and each
must be negative) i.e. ideal solutions are always completely miscible.
The equation above can be expressed in terms of chemical potentials of the individual
components

where

is the change in chemical potential of on mixing.

If the chemical potential of pure liquid is denoted


solution is

, then the chemical potential of in an ideal

Any component of an ideal solution obeys Raoult's Law over the entire composition range:

where
is the equilibrium vapor pressure of the pure component
is the mole fraction of the component in solution

It can also be shown that volumes are strictly additive for ideal solutions.

Non-ideality
Deviations from ideality can be described by the use of Margules functions or activity
coefficients. A single Margules parameter may be sufficient to describe the properties of the
solution if the deviations from ideality are modest; such solutions are termed regular.
In contrast to ideal solutions, where volumes are strictly additive and mixing is always complete,
the volume of a non-ideal solution is not, in general, the simple sum of the volumes of the
component pure liquids and solubility is not guaranteed over the whole composition range. By
measurement of densities thermodynamic activity of components can be determined.

Heat of Mixing

ENTHALPY CHANGES UPON MIXING


If seems to you that there are a lot of disconnected pieces of information in this section of material, you
should reinforce in your own mind the link between all of these pages. That link is simply that in order to
do energy balances we must be able to calculate changes in enthalpy associated with the process.

We always have the enthalpy of the outlet streams minus the enthalpy of the inlet streams
appearing in our equations. At this point we know several ways to get the enthalpy change
between inlet and outlet streams for specific situations. For example, we have seen how to use
steam tables to get this change when steam is the working fluid. We have used a PH diagram to
get enthalpies of a refrigerant. We have found that we can get changes in enthalpy associated
with sensible heat changes (changes in temperature) by integrating CpdT. And, on the last page,
we saw how the psychrometric chart can be used to get enthalpies for air + water systems.
But, what about mixtures? When two liquids are mixed, the final enthalpy is not necessarily the
sum of the pure component enthalpies. This is because the unlike interactions between molecules
is most likely different than the like interactions. Thus, if the A-B interactions are stronger than
the A-A and B-B interactions, then the mixing process will be exothermic (heat will be released
because the more tightly bound A-B interactions are at a lower energy). Again, what we want is
the change in enthalpy that occurs when the mixing occurs.
Let us examine the situation of the change in enthalpy upon mixing two liquid streams
isothermally as shown in the diagram at the right. n1 moles of liquid 1 are mixed with n2, and
the product liquid mixture leaves the mixer at the same temperature as the two inlet streams. We
will have to supply or remove a certain amount of heat, Q, in order to maintain the isothermal
process. A calorimeter is used to measure this quantity of heat that is given off or absorbed by
the mixing process. The energy balance for this
process can be written as,
Q = (n1 + n2) hm - n1h1 - n2h2
If we write this in terms of the heat
absorbed or released per mole of product solution, we obtain
Q/m = hm - x1h1 - x2h2 = hmix
This heat release or gain per mole of product solution for isothermal mixing is called
the heat of mixing, hmix. It may be either negative (exothermic - heat given off
because the mixture has a lower enthalpy than the pure components) or positive
(endothermic - heat absorbed because the mixture has a higher enthalpy than the
pure components). We can, of course, turn the equation around and write that for
an isothermal mixing process, the enthalpy of the mixture is the sum of the pure
component enthalpies (weighted by their mole fractions) plus the heat of mixing:

hm = x1h1 + x2h2 + hmix (isothermal)

Ideal Mixtures
An ideal mixture is one in which the interactions in the mixture are the same as for
the two pure components. Thus, for an ideal mixture,

hm = x1h1 + x2h2
hmix = 0

(ideal mixture)
(ideal mixture)

There are many mixtures that mix ideally or nearly ideally. All gases mix virtually
ideally, because the interactions are small between gas molecules. Even in liquids,
the interactions may be very similar between the components such that the heat of
mixing is small. For example, the heat of mixing of hexane and heptane would be
nearly zero. When ideal mixing can be assumed, one simply adds the contributions
from the pure components to obtain the enthalpy of the mixture.

Non ideal mixtures


As you can infer from what was said above, often you can tell from the types of
molecular interactions whether or not heats of mixing will contribute significantly to
the energy balance. Liquids that can associate, form hydrogen bonds, or even break
bonds as solvation occurs will often have fairly large heats of mixing. For example
the mixing of H2SO4 and H2O produces so much heat that you must be careful to add
the acid to the water very slowly so that it doesn't get hot enough to vaporize some
of the water.

In this class, we will ignore the heat of mixing term unless we are dealing with fairly large
quantities such as the solvation of acids and bases. Table B.11 on page 653 of your text lists heats
of mixing for HCl, NaOH, and H2SO4 with water. That same table is also reproduced and
available through the course tool links. Below is an example of how one can use these tabular
data in solving energy balance problems.

*******Using Heat of Mixing Data


Problem Statement:
What is the heat duty for a mixer that mixes 9.2 moles of H2O with 1 mole of 0.2 mole fraction
H2SO4 if the inlet and outlet streams are all to be at 25oC?
Solution:
The heats of mixing given in Table B.11 are recorded in terms of a variable r which is the moles
of water per mole of solute. We must first find this value of r (shown, however, as N in the figure
below) in order to obtain the value from the table.
The energy balance for this system is

Q = n3h3 - n1h1 - n2h2


We need to find the enthalpy of each
stream relative to a reference
enthalpy. We can do this because we
are always interested in a difference in
enthalpies and so the reference will
always cancel out. But, we must be sure that we choose the same reference for
each stream so that the cancellation really does occur. In this case, we choose the
pure components at 25oC as the reference enthalpy relative to which we will find the
values of all the streams.

Stream 1:
Stream 1 is a mixture of sulfuric acid and water at 25oC. It's enthalpy relative to the pure
references is simply the heat of mixing for this mixture since the pure enthalpies are identically
zero by definition of the reference: hm = x1h1 + x2h2 + hmix = 0 + 0 + hmix = hmix. Therefore,
we find r for this mixture in the following way, where A represents acid and W water:
xA = nA/(nW + nA) = 0.2 = nA/1 mol =====>
mol

nA = 0.2 mol

and

nW = 0.8

so r = nW /nA = (0.8 mol)/(0.2 mol) = 4.0


From Table B.11, we therefore obtain h1 = -54.06 kJ/mol A

Stream 2:
This stream is pure water which is the reference state. Therefore, relative to the reference state,
h2 = 0
Stream 3:
We now do a mass balance to find r for the outlet stream. The acid balance tells us
that 0.2 moles of A end up in the mixture and the water balance says that 9.2 + 0.8,
or 10 mol of water ends up in the final mixture. Thus,
r = nW /nA = (10 mol)/(0.2 mol) = 50
From Table B.11, we obtain h3 = -73.34 kJ/mol A

Now we complete the energy balance to obtain:


Q = (0.2 mol A)(-73.34 kJ/mol A + 54.06 kJ/mol A) = -3.9 kJ

Component Energy Balances on Reacting


Systems
Combining the energy and material balances for reacting systems
Consider the general energy balance for steady-state, steady-flow system (i.e.; CV = 0) with no
Ep and no EK between the inlet and outlet lines. In this case, the energy balance simplifies to:

where we are using the notation that we previously used for mass balances in
reacting systems. That is, the o superscript means inlet streams and no superscript
means the outlet or product streams. Note that the sums are over the components
not the streams. Here we consider only one inlet and one outlet stream, though the
results can be generalized to more streams if desired.

Recall the component material balance for reacting systems is: ni = nio + i. Let us substitute
this mass balance into the energy balance. There are two alternatives for doing this, we can
eliminate either the inlet or outlet number of moles in favor of the other. Let's look at both
alternatives.
Alternative 1: Replace outlet moles in energy balance with the material balance

Important: Note that the right hand side represents the change in enthalpy that occurs in
going from the inlet to the outlet stream via a specific path; namely,
reactants (Tin) ------------> reactants (Tout) --------------> products (Tout)
where during the first part of the path we heat up everything coming into the
reactor (hence the no term) to the outlet temperature (this is the sensible heat term
shown as the first term) and then during the second part of the path we react the
reactants to form the products. We know that we can choose any path that we want
in order to evaluate an enthalpy change, and this is just one particular path that
takes us from the reactants at Tin to the products at Tout.

Alternative 2: Replace inlet moles in energy balance with the material balance

Now replace the inlet moles using the material balance: nio = ni - i.

Important: Note that the right hand side represents the change in enthalpy that occurs in
going from the inlet to the outlet stream via a specific path; namely,
reactants (Tin) ------------> products (Tin) --------------> products (Tout)
Again, we have a two-step path that takes us from the inlet stream to the outlet
stream, and so the enthalpy change for this process represents the enthalpy out
minus the enthalpy in, which is what is required in the energy balance.
General Energy Balance for Reacting Systems
Let us generalize what we have above.

We can get the change in enthalpy from the inlet stream to the outlet
stream, as required by the DH in the energy balance, by calculating it
along any path that we want by using the heat of reaction and sensible
heat changes.
Obviously, the path that we will choose will be the one that makes the calculation
easiest for us. For example, we saw how to obtain heats of reaction at a reference
temperature (usually 25oC) from heats of formation, heats of combustion, or other
reactions. Therefore, often we will know the heat of reaction at the reference
temperature, Tref. It therefore makes sense to use a path that includes the reaction
occurring at Tref. In general, then, we can first change the temperature of the
components in the inlet stream to Tref, then perform the reaction, and then change
the temperature of everything in the product stream to the outlet temperature. This
process is: reactants (Tin) -------> reactants (Tref) ---------> products (Tref) -------->
products (Tout). The energy equation for this path is then:

Note that in this form, we can easily extend the energy balance to multiple
reactions. Instead of Hr, for multiple reactions we would have 1H1 + 2H2 + ...
for the reaction step.

Adiabatic Flame Temperature


Solving Energy Balances for Implicit Terms
So far, we have solved energy balances for Q or Ws. In reacting systems, it is
common to measure the heat and infer from this the extent of reaction or the outlet
condition. There is no new principle involved in doing this, but sometimes it makes
solution a little more complicated. Consider for example the case of finding the
temperature of the hottest part of a propane flame. If we try to find the maximum
possible temperature of that flame due to combustion of propane in air, we would
want to assume adiabatic conditions. This seems rather strange, because we have
all stood around a fire and felt the heat given off. However, if no heat were to be
lost, then we would have the highest possible temperature, so it is a useful
calculation to set an upper limit on the temperature of a flame. In practice, it
represents the hottest part of a flame better than what we might at first conjecture.
This is because the cooler parts of the flame will insulate the center, hottest part.

Most of the heat will be lost from the outer parts of the flame where the
temperature difference is greatest.

Consider the combustion reaction C3H8(g) + 5O2(g) = 3CO2(g) + 4H2O(v) where air is used as
the oxidant. To find the maximum possible flame temperature, assume Q = 0. We will feed the
fuel, propane, and air into the combustion unit at a temperature To. Now, regardless of the actual
pathway or mechanism for the combustion process, we can evaluate the process along a path of
our own choosing. Remember that all we need to do is find the H between the initial and final
state for any path since the value will be the same for all path between those two end points. We
will choose the path of combusting the fuel at To and then heating all of the products up to the
final temperature, T. This is diagrammed in the figure below.

In this figure, the actual path between the inlet and


outlet conditions might be represented by the hypotenuse of this triangle. We, instead, will use
the two-step pathway to get from the inlet to outlet conditions. The horizontal arrow represents
the reaction at constant temperature, and the vertical arrow represents the change in temperature
from To to T with no reaction.
The energy balance for this adiabatic process is
0 = Hout - Hin = Hr(To) + Hsensible

In these kinds of problems, we know the extent of reaction and the heat of reaction. We can then
use this equation to solve for the upper temperature limit in the integral over the heat capacities
that occurs in the sensible enthalpy change. This is best done in Mathcad because temperaturedependent heat capacities usually result in this being a fourth or fifth order equation. The
concrete example given below will help illustrate this.

Example of Adiabatic Flame Temperature Calculation

Problem Statement:
Calculate the adiabatic flame temperature of liquid butane burned with 30% excess
air. Both the air and liquid butane enter at 25 C.
o

Solution:
First we will perform the mass balance calculations so that we know the number of moles of
each component in the inlet and outlet streams.
Mass Balances
in: n O2 = (1 mol)(6.5)(1.3) = 8.45 mol
n N2 = (79/21)n O2 = 31.79 mol
o

note: 1.3 is the 30% excess O2

out: nO2 = 8.45 mol - (1 mol)(6.5) = 1.95 mol


nCO2 = (1 mol)(4) = 4 mol
nH2O = (1 mol)(5) = 5 mol
nN2 = n N2 = 31.79 mol
o

Energy Balances
Next we will set up the energy balance.
Q = Hout - Hin
We will calculate the change in enthalpy from the inlet to the outlet stream by using
a path of our choice. As was explained at the top of this page, the easiest path is to
react the materials at the inlet temperature and then heat everything in the outlet
stream up to the final temperature. Thus, we write,
Since adiabatic,

where Q has been set to zero since we are finding the maximum or adiabatic flame
temperature.

The heat of reaction at 25 C is easily found from the heat of combustion of propane. Note that it
is not exactly the heat of combustion because the water formed by the reaction is vapor whereas
o

it must be in the liquid phase for the heat of combustion to apply directly. By combining the
combustion reaction with the "reaction" of liquid water going to vapor so that the liquid terms
cancel, we find that the heat of reaction for this process is:
Hr = Hc + 5Hvap,H2O(25 C) = -2855.6 kJ/mol + 5(44.01 kJ/mol) = -2635.5 kJ/mol
o

The sensible enthalpy term is evaluated from the heat capacities of everything in
the outlet stream. Don't forget to include the nitrogen in the air and the excess
oxygen. All of these will be in the product stream and will absorb some of the heat
produced by the combustion process, thereby affecting the final computed adiabatic
flame temperature. Combining the heat of reaction term with the heat capacity
terms shown in the energy balance gives,

The heat capacities will be a function of temperature. The final temperature, the
upper limit for the integrals, will therefore be the only remaining unknown. We can
therefore solve for the adiabatic flame temperature using an appropriate solver. The
solution obtained in Mathcad is shown below.

Problem Statement:
Using Mathcad, find the adiabatic flame temperature of liquid butane when burned
with 30% excess air.

Solution:

Notes about these example


We can make some intereseting qualitative observations about the results of the above adiabatic
flame temperature calculations. For example, suppose we were to substitute pure O2 as the
oxidant in place of air, what would happen to the adiabatic flame temperature? From the energy
balance, we can see that N2 acts as an inert when air is fed and absorbs some of the heat of
combustion as sensible heat. In fact, a large fraction of the combustion gas effluent is nitrogen,
and so a large fraction of the sensible heat goes into heating the nitrogen up from 25oC to the
outlet temperature. Thus, a flame using pure O2 as the oxidant would be much hotter than one for
which air is the oxidizing stream.
If we perform the above calculations for a variety of molar ratios of air to fuel, we obtain a curve
like that at the right. Can you explain the general shape of this curve? Where do you think the
maximum occurs? Why does the temperature drop off as the number of moles of air is increased?

One final comment concerns terminology that is commonly


used when speaking about fuels. Fuels are often described in
terms of their heating value. In the previous example, H2O in
the product stream was vapor. Notice that we had to add the
heat of vaporization to the heat of combustion to get the heat
of reaction. Some of the released heat was used to vaporize
the water. The negative of this heat of reaction, when water is a vapor in the flue gas, is called the
lower heating value (LHV) of the fuel. The higher heating value (HHV) of a fuel is the negative
of the heat of combustion, i.e., H2O is a liquid in the flue gas. The difference between HHV and
LHV is the heat of vaporization of the water, multiplied by the appropriate stoichiometric
coefficient. Thus,

HHV = -Hc

LHV = -(Hc + waterHvap,water)

------ implies liquid H2O in flue gas


------ implies vapor H2O in flue gas

The LHV is the value that you would realize in an actual furnace or combustor where
water is a vapor in the hot flue gas.

Potrebbero piacerti anche