Thermo Final
Thermo Final
ENERGY
Energy is usually defined as the capacity to do work. Kinetic energy,
potential energy, chemical energy, and thermal energy are forms of energy
particularly useful to the chemist.
Kinetic energy is defined as the energy produced by a moving object.
For an object of mass m moving at a velocity v, the kinetic energy is equal
to ½ mv2.
Potential energy is energy available by virtue of an object’s position.
For instance, because of its altitude, a rock at the top of a cliff has more
potential energy and will make a bigger splash if it falls into the water 1
below than a similar rock located part way down the cliff.
Chemical energy can be considered a form of potential energy
because it is associated with the relative positions and the various attractive
and repulsive forces of atoms within a given substance.
Thermal energy is the energy associated with the random motion of
atoms and molecules. The thermal energy of a substance is related to its
temperature and quantity of matter in it. For example, the thermal energy
stored in a bathtub filled with water at 40 C is much more than the
thermal energy in a cup of coffee at 70 C.
HEAT
Heat can be defined as the transfer of thermal energy between two bodies
that are at different temperatures. Thus, we often speak of “heat flow” from
a hot object to a cold one, or “heat absorbed” when describing energy
changes that occur during a process.
1 cal = 4.184 J
1 L∙atm = 101.29 J
THERMODYNAMIC EQUILIBRIUM
A system is said to have attained a state of thermodynamic
equilibrium when it shows no further tendency to change its
properties with time.
The criterion for thermodynamic equilibrium requires that
three types of equilibrium exist simultaneously in a system. First
of all, the system must be in thermal equilibrium. This means
the temperature must be uniform throughout the system, and
also, the system must be at the same temperature as its
surroundings, if it is not insulated from the surroundings. Any
temperature differences that exist will cause heat to flow from a
higher temperature to a lower one until the temperature becomes
uniform throughout the system. Second, the system must be in
mechanical equilibrium. Mechanical equilibrium requires that
there be no macroscopic movement within the system itself, or from
the system to its surroundings. For example, consider a gas contained in a
cylinder equipped with a movable piston. If the piston is pushed in rapidly,
the rapid motion will set up pressure and temperature differences in the 3
gas, since the pressure and temperature in the region close to the piston
will be much greater than in a region away from the piston. Hence, no
single pressure measurement will unambiguously describe the system. If
enough time is allowed to elapse, these differences will level out until the
pressure and temperature throughout the gas will become uniform and
constant. The system is then in mechanical equilibrium. In a system
where more than one substance is present, thermodynamic equilibrium
requires that the composition of the system remains constant with time.
For chemical reactions this means no net chemical change can occur, i.e.,
the system must be at chemical equilibrium. Of course, this equilibrium
is a dynamic one in which the forward and reverse reactions are
continually occurring at equal rates. Hence there is no net change in the
concentration of reactants or products at equilibrium.
∆P = P2 – P1 (5-1)
The magnitude of P will be the same, whether the process was carried
out in one step or several steps, as long as P2 and P1 are equilibrium
values of the pressure for the two states. Not all variables of a system
possess this property of being independent of the paths taken between
states. The variables that do possess this property are called functions
or variables of state. They form an important class of variables that
arc very useful in thermodynamics. Pressure, temperature, and volume
are examples of state variables. Other very important state variables are
obtained from the laws of thermodynamics.
Eventually, of course, the metal ball must strike the ground and come to rest.
What happened then to the potential energy that the ball possessed origi-
nally? It is well known that this energy is converted into heat energy upon
striking the ground. This example illustrates that even though energy cannot
be created or destroyed, it can be converted into other forms. In the above
example mechanical energy was converted into heat energy. The First
law of thermodynamics is not concerned with how much of one form of
energy is converted into another. However, it requires that, when all the
energy changes have been accounted for, the total energy of the system
before and after the change must be the same.
∆E = Ef – Ei ( 5- 3)
E' = E i + q (5-4)
Now work is performed by the system, and this work must come from the
internal energy E' of the system. For instance, for a gas expanding against a
piston, the molecules of the gas perform work by pushing against the atmos-
phere. Because the expansion of a gas results in a decrease in the internal
energy, E’, of the system (Figure 5.2), work done by the system is assigned a
negative sign. This leads to a final state having E f given by
E f = E' + w (5-5)
Ef = Ei + q + w (5-6)
∆ E= E f – E i = q + w (5-
7)
It was mentioned above that the internal energy depends only on the
physical state in which the system exists. Thus ∆E in Eq. (5-3) does not de-
pend on the process or path by which the system went from the initial to the
final state and only depends on the initial and final states. This means that
the internal energy is another state function. This fact leads to a very
important conclusion, that it is impossible to construct a machine that will
produce work indefinitely without the input of energy, the so-called
perpetual motion machine. Consider a process which takes a system from
state A to a higher energy state B by path 1, as is shown in Figure 5.3, and
involves an input of heat q 1 , and work done by the system equal to w l . Let
us imagine another process (path 2 in Figure 5.3), involving an input of heat
q 2 and work done by the system equal to w 2 , which will take us back to 7
state A. The energy changes in these two processes are
∆ E 1 = E B – E A = q 1 + w1 path I (5-8a)
and
∆ E 2 = E A – E B = q 2 + w2 path 2 (5-8b)
The change in energy taken around the whole cycle ∆ E A → A , i.e., from
state A through state B and back to state A, is given by the sum of the Eqs.
(5-8a) and (5-8b),
q 1 + w 1 − w 2 –q 2 = 0
or
q 2 – q 1 = w1 – w2
The latter equation states that the net input of heat energy must just equal
the net work done by the system in a cyclic process. Only when this is true
will the cycle end up at a state with the same internal energy as it had orig-
inally. If it were possible to construct an engine that ran in a cycle in which
E 2 > E 1 , then one could continually produce work from the engine
without an equivalent expenditure of energy.
d w = f dr (5-11)
r2
w=−
∫Pr1
ext A dr (5-13)
V2
w=−
∫PV1
ext A dV (5-14)
w = −PextΔV (5-15)
EXAMPLE 5.1
A certain gas aexpands in volume from 2.0 L to 6.0 L at constant
temperature. Calculate the work done by the gas if it expands (a) against
vacuum ad (b) against a constant pressure of 1.2 atm.
Solution
(a) Since the external pressure is zero, the work w done is
w = −PΔV
= −0(6.0 − 2.0) L
=0
(b) The external, opposing pressure is 1.2 atm, so
w = −PΔV
= −4.8 L.atm
= −4.9 × 102 J
But V1 << Vg and Vl can be neglected when compared to Vv. In addition, if the
10
vapor is assumed to behave ideally, one has Vv = nRT/Pv, where n is the
number of moles of liquid that was converted to vapor and Pv is the equilib-
rium vapor pressure of the liquid at the temperature T. Using these
assumptions in Eq. (5-16), one obtains
nRT
w = − Pv Vv = − Pv = − nRT
(5-17)
Pv
This equation shows that for the work done in the vaporization of a
given amount of liquid depends only on the temperature and is independent
of the pressure or volume.
Note that Equation (5-17) also applies to reactions carried out in
open containers, which are accompanied by the production of gaseous
product. This is illustrated in example 5.3
EXAMPLE 5.2
Calculate the work done in vaporizing 1 mole of H 2O at 100°C assuming
ideality. 1 L∙atm = 101.3 J.
Solution
At 100°C, the normal boiling point, the vapor pressure of H2O is 1 atm. The
molar volume of H2O vapor at this temperature is given by
L 373 K
Vv = 22.4 × = 30.6 L / mol
mol 273 K
Thus the internal energy of the vapor is greater than that of the
liquid, as has been stated several times in this text, and this is due to the 11
large amounts of heat necessary to vaporize the liquid compared to the PV
work done by the vapor.
EXAMPLE 5.3
Calculate the work done when 50 g of iron dissolves in hydrochloric acid
in (a) a closed vessel and (b) an open beaker at 25°C. Treat the hydrogen
gas as perfect and ignore the initial volume of the system.
Solution
The reaction is
= 2.2 kJ
5.4 ENTHALPY
In most laboratory work in chemistry, reactions are carried out in open
flasks and therefore they occur at the constant pressure of the atmosphere.
For constant pressure processes in which pressure-volume work is the only
work done, Eq. (5-7) can be written, using Eq. (5-15) for w,
qp = ∆E + P ∆V (5-18)
where the subscript on q indicates that the heat change occurs at constant
pressure. If, however, the process is carried out at constant volume, ∆V is
zero and one obtains
qv = ∆E (5-
19)
i.e., the change in internal energy for a constant volume process, in which
no PV work is done, is equal to the heat absorbed or evolved by the sys-
tem. Because of the importance of constant pressure processes, it is con-
venient to define a new thermodynamic function, the enthalpy, given the
12
symbol H. The enthalpy is defined by the equation
H = E + PV (5-20)
Since E, P, and V are all state functions, it follows that H is also a state
function and is therefore, independent of the path taken between initial and
final states of the system. One can write, from Eq. (5-18), for the change in
enthalpy ∆H,
∆ H = ∆ E + ∆ (PV) (5-26)
∆ H = ∆ E + P ∆ V+ V ∆ P (5-27)
∆H = ∆E + P∆V (5-28)
Thus one sees that the change in enthalpy for a constant pressure process,
in which the only work done is PV work, is equal to the heat absorbed or
evolved in the process. Since H = H f – H i , where H f and H i are
the enthalpies of the final and initial states, respectively, then when
heat is absorbed by the system, the positive value for H means that the
heat content of the final state is greater than that of the initial state. When
heat is evolved by the system, H has a negative sign and the heat content
of the final state is less than that of the initial state. Using qp and qv for
H and E, respectively, in Eq. (5-28), one obtains
qp = qv + P∆V (5-24)
ΔH
Cp = (5-26b)
ΔT
The first term in the expression on the right is equal to Cv by Eq. (5-
26a). The second term is simply equal to R for one mole of an ideal gas.
(Show this.) Hence, one has
C p = Cv + R (5-28)
Cv (J/mol·K) Cp (J/mol·K)
H e , N e , A r, K r, X e
12.5 20.78
H2
20.53 28.84
N2
21.06 29.37
CO2
28.81 37.12
H2O(g)
16.46 24.77
NH3
27.32 35.63
CH4
27.33 35.64
5.7 TH ER M O C H E M I ST R Y
Up to now, consideration has been given only to energy changes in -
volved in physical processes. One of the most important applications of 14
the First law to chemistry is in the study of the heat changes that
occur in chemical reactions. This study forms the subject of
thermochemistry.
It was shown in the above discussions that ∆ E and ∆H were
independent of the path taken in going from the initial to the final state
of a system. Thus for reactants going to products in a chemical
reaction, the only important consideration is the measurable
quantities q v (= ∆ E) and q p (= ∆ H), the heat changes involved in the
reaction occurring at constant volume and constant pressure,
respectively. The absolute values of the internal energies and
enthalpies of reactants and products are not required, since one is in-
terested only in the change in these values for the reaction. When
heat is evolved in a reaction, the final state (products) is lower in
energy and by the convention used before, q is negative. For a
constant pressure reaction (for instance, a reaction run in an open
beaker), qp , and. therefore. ∆H, is negative. This fact is indicated by
writing the reaction
Reactants → products ∆ H T = –q J
where q is the heat evolved in the reaction at the absolute or Kelvin
temperature T. For example, the reaction of sulfur and oxygen gas to give
sulfur dioxide at 25°C is found to evolve 296.9 kJ for 1 mole of SO 2
formed. This is written
ENTHALPY OF COMBUSTION
The heat involved due to the complete combustion of 1 mole of a
hydrocarbon with oxygen is known as the enthalpy of combustion. As
an example, one mole of propane, C 3 H 8 (g), is burned in O 2 (g) at 1 atm
and 25°C to give CO 2 (g) and H2O(l), and the heat evolved is found to be
2220 kJ/mol. Since all substances are at 1 atm and 25°C, they are in their
standard states, and the heat evolved is the standard enthalpy of
combustion. This change is expressed in an equation as
Hydrocarbon H (kJ/mol)
Methane, CH4(g) –890.36
Ethane, C2H6(g) –1559.88
Propane, C 3 H 8(g) –2220.07
n-Butane, C4H 1 0 (g) –2878.51
Isobutane. C4H10)(g) –2871.65
Ethylene, C2H 4(g) –1410.97
Acetylene, C2H 2 (g) –1299.63
Benzene, C 6H 6(l) –3267.62
16
5.10 HESS’S LAW
In 1840, Hess put forth the following empirical law: the overall heat
change at constant pressure or constant volume in a given chemical reaction
is the same regardless of whether it takes place in one direct step or in sev-
eral steps. Thus if compound B could be formed from A directly and it were
also possible to go from A to B through intermediates C and D, the enthalpy
change for the direct path would be the same as the sum of the enthalpy
changes involved in the intermediate reactions. This is shown schematically
in Figure 5.5, where the ∆H's are the enthalpy changes for the individual
reactions proceeding in the directions shown by the arrows. Hess's law re-
quires that ∆H 4 = ∆H1 + ∆H 2 + ∆H3 . It can be shown that this result can
be derived from the First law. Thus since ∆H is a state function, the sum of
the ∆H's around the cycle, taken from A through C, D and B and back to A,
by the First law must be zero [see Eq. (5-10)]. This means
∑ ΔH i = 0
= ∆H1 + ∆H2 + ∆H3 + (–∆H4) (5-29)
i
or
which is the same result predicted by Hess's law. In Eq. (5-29) the
sign of ∆ H 4 was reversed since in going around the cycle in a
counterclockwise manner, the reverse of the reaction path A → B was
traversed. If heat were evolved in going from A → B, the First law
requires that the same amount of heat be absorbed in the reverse
reaction. Thus ΔH is equal in magnitude but opposite in sign for the reverse
reaction.
The importance of Hess's law lies in the fact that the enthalpy
changes can be calculated for reactions that cannot be performed in the
laboratory. In order to obtain reliable H values for reactions, the reactions
must proceed rapidly and must go to completion. Many reactions are too
slow or involve the formation of side products and therefore do not meet 17
these criteria. Hess's law can be used to obtain H for these reactions.
For example, it is not possible to measure directly the heat evolved when
graphite is reacted with O2 to give CO, since it is impossible to prevent the
formation of CO2. However, the heat evolved in the complete combustion
of graphite to CO2 can be measured, as well as the heat evolved in the
combustion of CO to CO2. At 18°C, these reactions are
from the measured values of H for the above two reactions, reverse
reaction (2) and add to reaction (1). The same procedure must be followed
with the enthalpy changes for these reactions. Reversing the direction of
reaction (2) means the sign of Hrxn2 must be reversed, i.e., the reaction
CO2(g) → CO(g) + ½O2(g) requires the input of 282.9 kJ. The result then is
ENTHALPY OF FORMATION
When a mole of a given compound is formed in its standard state from its
elements, all in their standard states, the heat evolved or absorbed is known
as the standard enthalpy of formation of the compound, ΔH f . As an
example, the formation of SO 2 (g) from its elements at 25°C and 1
atm evolves –296.90 kJ/mol of SO 2 (g) formed, and, therefore, ΔH f = –
296.90 kJ/mol at 25°C and 1 atm.
S(rhombic) + O2(g) → SO 2 (g) ΔH f = –296.90 kJ/mol
aA + bB + . . . → cC + dD + . . .
if ΔH f for products and reactants are known, ∆H for the reaction can be 18
found from Hess's law in the form
∑ n i ( ΔH )i ∑ nj ( ΔH ) j
ΔH rxn = f (products) — j f (reactants) ( 5-30)
i
where ni are the coefficients of the products and nj are coefficients of the re-
actants, in the balanced chemical equation.
Now one can see why it is convenient to choose the enthalpy of the
elements in their standard states as zero. Since H for the process is the
difference in enthalpies of the product (final state) and reactants (initial
state), then for the formation of one mole of methane, CH 4, under standard
state conditions, the measured value of the enthalpy change for the reaction
becomes
ΔHrxn = (1 mol) ΔH f ( CH4 , g ) – (1 mol) ΔH f ( C, graphite) – (1 mol) ΔH f ( H2 , g)
Substance ΔH f (kJ/mol)
H 2 0(g) –241.8
H 2 0(l) –285.8
HCI(g) –92.3
S0 2 (g) –296.1
S O 3 ( g) –395.2
NH 3(g) –46.3
CO(g) –110.5
CO 2 (g) –393.5
CH4 (g) –74.7
C 2 H 2 (g) +226.6
NO(g) +90.4
NO 2 (g) +33.85
3H2(g) + 6C(graphite) → C6 H6 ( l) ΔH f = ?
19
The enthalpy change in the written reaction cannot be determined
directly since the reaction of carbon and hydrogen results in a mixture of
hydrocarbons. However, one can obtain ΔH f of benzene by noting that
benzene, like all organic compounds containing carbon and hydrogen can
be burned in oxygen to produce H2O and CO2, whose ΔH f ’s are known.
C6H6(l) + 15 O2(g) → 6 CO2(g) + 3 H2O(l) ΔH rxn = –3267.6 kJ
– (15 mol) ΔH f ( O2 , g)
ΔH rxn .
EXAMPLE 5.4
Calculate the enthalpy of reaction at 25°C for the following reaction from
the standard heats of formation of reactants and products.
EXAMPLE 5.5
Pentaborane-9, B5H9, is a colorless, highly reactive liquid that was
considered as a potential rocket fuel because it produces a large amount of
heat per gram when it explosively burns in oxygen. The idea was abandoned
because the solid B2O3 formed by the combustion is abrasive and would
quickly destroy the nozzle. The reaction is
20
2 B5H9(l) + 12 O2(g) → 5 B2O3(s) + 9 H2O(l)
The standard heat of formation at 25°C for B5H9, B2O3, and H2O are
respectively, 73.2 kJ/mol, −1263.6 kJ/mol, and −285.8 kJ/mol.
Calculate the heat liberated per gram of B5H9.
Solution
Using the given ΔH f values, we write,
ΔH rxn = (5 mol) ΔH f ( B2O3,s) + (9 mol) ΔH f (H2O,l) (2 mol) ΔH f (B5H9,l) − 0
= −9036.6 kJ
= −71.58 kJ/g B5 H9
Note that compounds with positive ΔH f values release more heat during
combustion and are less stable than those with negative ΔH f values.
∆H = ∆E + P∆V
one can see that if there is no net change in volume during a reaction, ∆V =
0 and ∆ H = ∆ E. Two main types of reactions correspond to this
condition. One type includes reactions involving only liquids and solids. In
this case, ∆V is negligibly small and P∆V is usually negligible in comparison
to ∆E. The second type includes reactions in which the number of moles of
gaseous products equals the number of moles of gaseous reactant, thus
resulting in no net increase in volume, and ∆V is essentially zero.
Let us consider the case where there is a change in the number of
moles of gaseous substances during the reaction. In this case P∆V will be a
significant quantity and must be taken into account in Eq. (5-22).
Assuming the gases to behave ideally, one can write for P∆V
P ∆ V = ∆ n RT (5-31)
This equation gives the relationship between the heats of reaction at constant
pressure and at constant volume for a reaction involving PV work only.
EXAMPLE 5.6
Calculate the heat change for the combustion of n-pentane when the
reaction is run in a Paar bomb at 25°C. The reaction is
C 5 H 1 2 ( l ) + 8 O 2 ( g ) → 5 CO2(g) + 6 H2O(l) ΔH rxn = –3536. 2 kJ
Solution
The value of ΔH rxn is –3536. 2 kJ. The value of ∆n is (5 – 8) = –3 moles,
and using Eq. (5-32), one has for the constant volume reaction
∆E = ∆H –∆nRT
= – 3.5362 × 10 6 J – ( – 3 mol × 8.314 J / mol ⋅ K × 298 K )
= –3. 53 J
and the enthalpy change for the process is H a t o m . For this particular
molecule, H a t o m corresponds to the total amount of energy needed to
break all the C—H bonds in one mole of CH 4 ; therefore, division of
H a t o m by 4 would give the average C—H bond energy in methane,
expressed in kJ/mol (or kcal/mol).
4 H(g) + C(g)
1 2 3
Figure 5.6 shows how we can use the standard heat of formation,
ΔH f to calculate the atomization energy. Across the bottom we have the
chemical equation for the formation of CH4 from its elements. The enthalpy
change for this reaction, of course, is ΔH f . In this figure we also can see an
alternative three-step path that leads to CH4(g). One step is the breaking of
H—H bonds in the H2 molecules to give gaseous hydrogen atoms, another is
the vaporization of carbon to give gaseous carbon atoms, and the third is
the combination of the gaseous atoms to form CH 4 molecules. These
changes are labeled 1, 2, and 3 in the Figure.
Since H is a state function, the net enthalpy change from one state
to another is the same regardless of the path that we follow. This means
that the sum of the enthalpy changes along the upper path must be the
same as the enthalpy change along the lower path, ΔH f . Perhaps this can
be more easily seen in Hess's-law terms if we write the changes along
the upper path in the form of thermochemical equations.
Steps 1 and 2 have enthalpy changes that are called standard
heats of formation of gaseous atoms. Values for these quantities
have been measured for many of the elements and some of them are
given in Table 5.5. Note that for diatomic gaseous molecules these
values are exactly half the bond energy in the molecule. Step 3 is the
opposite of atomization, and its enthalpy change will therefore be the
negative of H a t o m ( r ecal l t hat i f w e r ever se a r eact i on, w e change
t he si gn of i t s H ) .
Let’s substitute for ΔH1 , ΔH 2 , and ΔH 3 , and then solve for Hatom. First,
we substitute for the ΔH quantities.
Changing sign and rearranging the right side of the equation gives
Now all we need are values for the ΔH f ’s on the right side. From Table 5.5
we obtain ΔH f [H(g)] and ΔH f [C(g)], and the value of ΔH f [CH4(g)] as
already seen is equal to 74.8 kJ/mol obtained from 5.3.
ΔH f [H(g)] = 218.0 kJ/mol
and division by 4 gives an estimate of the average C—H bond energy in this
molecule.
1662 kJ / mol
Bond energy = = 415.5 kJ / mol C − H bond
4
This value is quite close to the value in Table 5.6, which is an average of C
—H bond energies in many different compounds. The other bond energies in
Table 5.6 are also based on thermochemical data and were obtained by
similar calculations.
EXAMPLE 5.7
Use the bond energies in Table 5.5 and 5.6 to estimate the heat of formation
of methanol vapor, CH3OH(g).
Solution
We solve this problem in much the same way that we calculated the
bond energy of methane in the discussion above. We setup two paths from
the elements to the compound, as shown in the Figure shown below. The
lower path has an enthalpy change corresponding to ΔH f [CH3OH(g)], while
the upper path takes us to the gaseous elements and then through the 25
energy released when the bonds in the molecule are formed. This latter energy
can be computed from the bond energies in Table 5.6. As before, the sum of
the energy changes along the upper path must be the same as the energy
change along the lower path, and this permits us to compute ΔH f [CH3OH(g)].
1 2 3 4
Steps 1, 2, and 3 involve the , formation of the gaseous atoms from the
elements, and their enthalpy changes are taken from Table 5.5.
ΔH1 = 1 mol × ΔH f [C(g)] = 1 mol × 715.0 kJ/mol = 715.0 kJ
ΔH 2 = = 4 mol × ΔH f [H(g)] = 4 mol × (218.0 kJ/mol) = 872.0 kJ
Adding these values gives a total energy input for the first three steps of 1836.2
kJ. In other words, the net ΔH for the first three steps is +1836.2 kJ.
The formation of the CH3OH molecule from the gaseous atoms is
exothermic—energy is always released when atoms become joined by a
covalent bond. In this molecule we can count three C—H bonds, one C — O
bond, and one O—H bond. Their formation releases energy equal to their
bond energies, which we obtain from Table 5.6.
Bond Energy
(kJ)
3(C—H) 3 × (413 kJ/mol) = 1239
C —O 351
O—H 464
Adding these together gives a total of 2054 kJ. ΔH for this step is therefore
—2054 kJ (because it is exothermic). Now we can compute the total
enthalpy change for the upper path.
ΔH = –218 kJ
The value just calculated must be equal to the ΔH for the lower path
of the Figure, which is ΔH f for CH3OH(g). Experimentally, it has been found
that ΔH f for this molecule (in the vapor state) is –201 kJ/mol. At first glance,
the agreement doesn't seem very good, but on a relative basis the calculated
value (–218 kJ) differs from the experimental one by only about 8%.
26
Comparisons between measured and calculated bond energies
have sometimes helped chemists understand unusual properties of
some substances. Consider, for example, the case of benzene, C 6 H 6 .
This molecule has its six carbon atoms arranged in a hexagonal ring,
and bonded to each carbon is one of the hydrogen atoms. One of the
Lewis structures that we can draw for the molecule is
C p = a + bT + c T 2 + dT 3 (5-33a)
or
C p = a + bT + cT – 2 ( 5- 33b)
and substituting the expression for Cp from Eq. (5-33a and 5-33b) into Eq. (5-
34), one obtains an expression for the change in enthalpy, dH, of a
substance in some phase for a change in temperature dT
dH = ( a + bT + cT 2 + dT 3 ) dT (5-35a)
or
dH = ( a + bT + cT – 2 ) dT ( 5- 35b)
and
ΔH = a (T 2 – T1 ) +
2
(
b 2
) c
3
(d
T 2 – T12 + T23 – T13 + T24 – T14
4
) ( ) (5-
37)
or
T2
ΔH = H T2 – H T1 =
∫
T1
(a + bT + cT –2 ) dT (5-38)
and
ΔH = a(T2 – T1 ) +
b 2
2
( 1
T2 – T12 − c
T
–)1
T
(5-39)
2 1
If either the temperature change or the change in heat capacity over the tem-
perature range is small, i.e., b and c in Eq. (5-33a or 5-33b) are small, then 28
Cp can be taken equal to the constant value a. In this case, Cp is independent
of temperature, and integration of Eq. (5-36 or 5-38) gives
EXAMPLE 5.8
Calculate the heat necessary to raise the temperature of 2 moles of gaseous
C12 from 500 K to 1000 K at constant pressure.
Solution
The expression for Cl2 of Cl2 as a function of temperature is obtained from
Table 5.7. Using this in Eq. (5-36), one has for one mole of gas
1000
ΔH = q p =
∫Cp dT
500
1000
=
∫ (37.03 + 0.67 × 10
500
–3
T –2.85 × 10 5 T –2 ) dT
aA + bB → cC + dD
At T1, the enthalpy change is measured to be ∆H1. One now would like to
know the enthalpy change at temperature T2, where T2 > T1. This result can
be obtained by considering the cyclic process shown in Figure 5.7.
30
∆H’ ∆H”
where
Cp(reactants) = aC p (A) + bC p (B)
where
Cp(products) = cCp(C) + dCp(D)
We can change the sign of the second term on the right-hand side of this
equation if we reverse the limits of integration. Thus
T2 T2
ΔH 2 = ΔH1 +
∫
T1
C p(products) dT –
∫C
T1
p(reactant s) dT (5-42)
31
The expression can be made more compact if we define
T2
ΔH 2 = ΔH1 +
∫ [(cC (C) + dC (d) ) – (aC
T1
p p p p(A) )]
+ bCp(B) dT (5-44)
EXAMPLE 5.9
The hydrogenation of methyl chloride was studied at 248°C (521 K). It was
found that at that temperature ∆H521 = –82.32 kJ. Calculate ΔH 298 by
using Cp values reported at 400 K. Cp(CH4(g)) = 41.28 J/mol·K, Cp(HCl(g) =
29.11 J/mol·K, Cp(CH3Cl(g) = 48.46 J/mol·K, and Cp(H2(g) = 28.56
J/mol·K.
Solution
ΔH 298 is related to ΔH 521 by Eq. (5-45)
ΔH 298 = ΔH 521 + Cp T
and so
ΔH 298 = ΔH 521 + Cp(298 – 521)
= –80.84 kJ
EXAMPLE 5.10
One method for producing ethanol on an industrial scale is the direct
hydration of ethylene
C2H4(g) + H2O(g) → C2H5OH(g) ΔH 298 = –42.92 kJ
.
Calculate ΔH 50
Solution
The relationship between ΔH 298 is given by Eq. (5-43), (5-44)
and ΔH 50
50
ΔH 50
= ΔH 298 + ∫ ΔC dT
298
p
= ∆a + ∆bT + ∆c T 2 + ∆d T 3
∫298
ΔC pdT =
298
–2
T) –(3.088 × 10 –5 T 2 ) + (5.979 × 10 –9 T 3 ) dT
33
= (–16.29)(50 – 298) +
2
(
(5.13 × 10 – 2 )
)
(50)2 – (298)2
–
3
(
(3.088 × 10 – 5 T 2 )
)
(50)3 – (298 )3 + (
(5.979 × 10 – 9 )
4
(50)4 – (298)4 )
= 4039.92 – 2211,54 + 271.11 – 11.78
= 2087.71
= ΔH
ΔH 50 298 + 2087.71
5.17 ENTROPY
In order to predict the spontaneity of a process, we need to know two things
about the system. One is the change in enthalpy, which is nearly
equivalent to E for most processes. The other is entropy (S), which is
a measure of the randomness or disorder of a system. The greater the
disorder of a system, the greater its entropy. Conversely, the more ordered a
system, the smaller its entropy. One way to illustrate order and disorder is
with a deck of playing cards. A new deck of cards is arranged in an
ordered fashion (the cards are from ace to king and the suits are in the
order of spades to hearts to diamonds to clubs). Once the deck has been
shuffled, the cards are no longer in sequence by number or by suit. It is
possible, but extremely unlikely, that by reshuffling the cards we can
restore the original order. There are many ways for the cards to be Out
of sequence but only one way for them to he ordered according to our
definition.
For any substance, the particles in the solid state are more
ordered than those in the liquid state, which in turn are more ordered
than those in the gaseous state. So for the same molar amount of a
substance, we can write
∆S = Sf – Si (5-46)
where Sf and Si are the entropies of the system in the final and
initial states, respectively. If the change results in an increase in
randomness, or disorder, then Sf > Si or ∆S > 0. Thus, both melting
and vaporization processes have ∆S > 0. The solution process usually
leads to an increase in entropy. The following example deals with the 37
entropy changes of a system resulting from physical changes.
EXAMPLE 5.11
Predict whether the entropy change is greater than or less than
zero for each of the following processes: (a) freezing ethanol, (b)
evaporating a beaker of liquid bromine at room temperature, (c)
dissolving sucrose in water, (d) cooling nitrogen gas from 80°C to
20°C.
Answer
(a) This is a liquid-to-solid phase transition. The system becomes
more ordered, so that ∆S < 0.
(b) This is a liquid-to-vapor phase transition. The system
becomes more disordered, and ∆S > 0.
(c) A solution is invariably more disordered than its components (the
solute and solvent). Therefore, ∆S > 0.
(d) Cooling decreases molecular motion; therefore, ∆S < 0.
PRACTICE EXERCISE
How does the entropy of a system change for each of the following
processes? (a) condensing water vapor, (b) forming sucrose
crystals from a supersaturated solution, (c) heating hydrogen gas
from 60°C to 80°C, (d) subliming dry ice
For a spontaneous process, the second law says that ∆Suniv must be
greater than zero, but it does not place a restriction on either ∆Ssys or
∆Ss u r r . Thus it is possible for either ∆Ssys or ∆Ss ur r to be negative, as
long as the sum of these two quantities is greater than zero. For an
equilibrium process, ∆Suniv is zero. In this case ∆Ssys and ∆Ss ur r must be
equal in magnitude, but opposite in sign. What if for some process we
find that ∆Suniv is negative'? What this means is that the process is not
spontaneous in the direction described. Rather, it is spontaneous in the
opposite direction.
aA + bB → cC + dD
ΔS rxn = [cS (C) + dS (D)] – [aS (A) + bS (B) (5-
49)
ΔS rxn = Σn S (productd) – ΣmS (reactants) (5-
50)
EXAMPLE 5.12
From the absolute entropy values the chemicals taking part in the
reactions given below, calculate the standard entropy changes for the
following reactions at 25°C. S° (CaO) = 39.8 J/K∙mol, S°(C02) =
213.6J/K∙mol, S°(CaCO3) = 92.9 J/K∙mol, S° (NH3) =193 J/K∙mol, S°(N2)
=192 J/K∙mol, S°(H2) = 131 J/K∙mol, S°(HCl) = 187 J/K∙mol, S°(Cl2) = 223
J/K∙mol.
Answer
We can calculate ∆S° by using Equation (5-60).
(a) ΔS rxn = [S° (CaO) + S° (C02)] – [S° (CaCO3)]
= [(1 mol)(39.8 J/K∙mol) + (1 mol)(213.6J/K∙mol)]
– (1 mol)(92.9 J/K∙mol)
= 160.5 J/K
= –199 J/K
= 20 J/K
EXAMPLE 5.14
Predict whether the entropy change of the system in each of the
following reactions is positive or negative.
(a) 2 H 2 (g) + O 2 (g) → 2 H2O(l)
(b) NH 4 Cl(s) → NH 3(g) + HCl(g)
(c) H 2 (g) + Br 2 (g) → 2 HBr(g)
Answer
(a) Two gases combine to form a liquid. Therefore, S is negative.
(b) Since the solid is converted to two gaseous products, S is positive.
(c) We see that the same number of moles of gas is involved in the
reactants as in the product. Therefore we cannot predict the sign of S,
but we know the change must be quite small.
PRACTICE EXERCISE
Discuss qualitatively the sign of the entropy change expected for each of the
following processes:
(a) I 2 (g) → 2 I(g)
(b) 2 Zn( s ) + 0 2 (g) → 2 ZnO(s)
(c) N 2 (g) + O 2 (g) → 2 NO(g) 40
ENTROPY CHANGES IN THE SURROUNDINGS
Next we see how the ∆Ss u r r is calculated. When an exothermic process
takes place in the system, the heat transferred to the surroundings
enhances motion of the molecules in the surroundings.
Consequently, there is an increase in disorder at the molecular level,
and the entropy of the surroundings increases. Conversely, an
endothermic process in the system absorbs heat from the
surroundings and so decreases the entropy of the surroundings
because molecular motion decreases. For constant-pressure
processes the heat change is equal to the enthalpy change of the
system, ∆Ssys . Therefore, the change in entropy of the surroundings, ∆Ss u r r
is proportional to
∆Ss u r r ∝ – ∆Ssys
The minus sign is used because if the process is exothermic, ∆Ssys
is negative and ∆Ss ur r is a positive quantity, indicating an increase in
entropy. On the other hand, for an endothermic process, ∆Ssys is
positive and the negative sign ensures that the entropy of the
surroundings decreases.
The change in entropy for a given amount of heat also depends on the
temperature. If the temperature of the surroundings is high, the
molecules are already quite energetic. Therefore, the absorption of heat
from an exothermic process in the system will have relatively little
impact on molecular motion and the resulting increase in entropy will be
small. However, if the temperature of the surroundings is low, then the
addition of the same amount of heat will cause a more drastic increase in
molecular motion and hence a larger increase in entropy. By analogy,
someone coughing in a crowded restaurant will not disturb too many
people, but someone coughing in a library definitely will. From the
inverse relationship between ∆Ssurr and temperature T (in Kelvins)—that
is, the higher the temperature, the smaller the ∆Ssurr and vice versa—we
can rewrite the above relationship as
– ΔH sys
ΔS surr = (5-61)
T
Let us now apply the procedure for calculating ∆Ss ys and ∆Ss u r r to
the synthesis of ammonia and ask whether the reaction is spontaneous at
25°C:
∆Ss ys = – ∆ Ss ur r
– ΔH sys ΔH sys
ΔS sys = – =
(5-52)
T T
The change in entropy as a result of vaporization, fusion, and
sublimation can be written as
where ∆Hfus, ∆Hvap, and ∆Hsub are the heats of fusion, vaporization, and
sublimation, respectively, and Tfus, T boil, and T sub are the fusion, boiling,
and sublimation temperatures, respectively.
As an example, let us first consider the ice-water equilibrium. For
the ice → water transition, ∆Hsys is the molar heat of fusion, equal to
6010 J/mol and T is the normal melting point. The entropy change is
therefore
6010 J / mol
= 22.0 J / K ⋅ mol
273 K
– 6010 J / mol
= – 22.0 J / K ⋅ mol
273 K
42
In the laboratory we normally carry out unidirectional phase
changes, that is, either ice to water or water to ice. We can calculate
entropy change in each case using the equation ∆S = ∆H/T as long as the
temperature remains at 0°C. The same procedure can be applied to the
water-steam transition. In this case ∆H is the heat of vaporization and T is
the boiling point of water. Example 5.14 examines the phase transitions in
benzene.
EXAMPLE 5.14
The molar heats of fusion and vaporization of benzene are 10.9 kJ/mol and
31.0 kJ/mol, respectively. Calculate the entropy changes for the solid
liquid and liquid → vapor transitions for benzene. At 1 atm pressure,
benzene melts at 5.5°C and boils at 80.1°C.
Solution
At the melting point, the system is at equilibrium. Therefore ∆S, the
entropy of fusion is given by
ΔH fus
ΔS fus =
T fus
T2 T2
∫ ∫
dH Cp
ΔSsys = = dT (5-54)
T1 T T1 T
T2
ΔS = C p ln 43
T1
If the system is heated very slowly the change in the surroundings
is equal and opposite in sign to that for the system, and
S s ys + Ssurr = Suniv = 0
T2 T2
∫ ∫
dE Cv T
ΔSsys = = dT = Cv ln 2 (5-65)
T1 T T1 T T1
Again for a reversible change, the entropy change for the system plus
surroundings is zero.
= Sf
∫
Cp
ΔS = S f = dT (5-56)
0 T
Firure 5.8
T2 Cp T2 (a + bT + cT ) dT = a ln T + bT −
–2
2c 2
T
ΔS = S = ∫T
1 T
dT = ∫T1 T
T –2 T1
T1 C T1 T1
kT3
∫ ∫ ∫
p(s)
ΔS = dT = dT = k T 2 dT
0 T 0 T 0
EXAMPLE 5.15
The molar Cp of a solid at 10 K is 0.43 J/K∙mol. What is the entropy of the
solid at that temperature?
Solution
Cp at 10 K is described by the Debye Equation
C p = kT 3 J / K ⋅ mol –1
T Cp T T3 T
∫0 ∫0 ∫0 T
2
S = S (0) + dT = 0 + k d T = k dT
T T
S=
1
3
[
k T 3 − (0 K )3 ]
Substituting
S=
1
3
[
( 4.3 × 10 − 4 J / K ⋅ mol ) (10 K )3 − (0 K )3 ]
S = 0.14 J/K∙mol
∫
p
ΔS2 = dT = 93.4 J ⋅ K –1 ⋅ mol –1
12.04 T
∫
C p'
ΔS4 = dT = 15.4 J ⋅ K –1 ⋅ mol –1
220.4 T
∫
C p"
ΔS6 = dT = 10.7 J ⋅ K –1 ⋅ mol –1
264.5 T
This equation says that for a process carried out at temperature T, if the
changes in enthalpy and entropy of the system are such that ∆H sys – T ∆
Ssys is less than zero, the process must be spontaneous.
In order to express the spontaneity of a reaction more directly, we
can use another thermodynamic function called Gibbs free energy (G), or
simply free energy:
G = H – TS (5-60)
∆G = ∆H – T∆S (5-61)
To calculate ΔGrxn we start with the equation
aA + bB → cC + dD
or, in general,
ΔGrxn = Σ n ΔG f ( p r o d u c t s ) – Σ n ΔG f ( r e a c t a n t s ) (5-62)
EXAMPLE 5.16
Calculate the standard free-energy changes for the following reactions at
25°C.
(a) CH 4 (g) + 2 O 2 (g) → CO2(g) + 2 H2O(l)
(b) 2 MgO(s) → 2 Mg(s) + O2(g)
Solution
(a) According to Equation (5-62), we write
ΔGrxn = [(1 mol) ΔG f (CO2) + (2 mol) ΔG f (H2O) – [(1 mol) ΔG f (CH4)
+ (2 mol) ΔG f (O 2 )]
= 1139 kJ
In the preceding example the large negative value of ΔGrxn for the
combustion of methane in (a) means that the reaction is a spontaneous
process under standard-state conditions, whereas the decomposition of
MgO in (b) is nonspontaneous because ΔGrxn is a large, positive quantity.
Remember, however, that a large, negative ΔGrxn does not tell us
anything about the actual rate of the spontaneous process; a mixture
of CH 4 and O2 at 25°C could sit unchanged for quite some time in the
absence of a spark or flame.
H S G Example
= 177.8 kJ
0 = ∆H° – T∆S°
ΔH
T =
ΔS
(177.8 kJ) (1000 J /1 kJ)
T = = 1108 K or 835°C
160.5 J / K
∆G ° = ∆H° – T ∆S°
= 177.8 kJ – (1113 K)(160.5 J/K)(1kJ/1000 J)
= –0.8 kJ
PHASE TRANSITIONS
At the temperature at which a phase transition occurs (the melting point or
boiling point), ∆G = 0 so Equation (5-61) becomes
0 = ∆H – T∆S
ΔH
ΔS =
T
where ∆H is ∆Hfus, ∆Hvap, or ∆Hsub; the heat of fusion, vaporization, or
sublimation, and T is T fus, T boil, or T sub; the fusion, boiling, or sublimation
temperature.
Notice that the above equation is the same as Equation (5-53) and
can be applied to calculate entropy changes during phase transitions.
(See Example 5.14).
∆G = ∆G° + RT l n Q (5-63)
Case 1: If ∆G° is a large negative value, the RT l n Q term will not become
positive enough to match the ∆G° term until a significant amount of
product has formed.
or
EXAMPLE 5.17
Using data listed in an Appendix, calculate the equilibrium constant
(K p ) for the following reaction at 25°C:
Solution
According to Equation (5-62)
ΔGrxn = [2 ΔG f (H 2 ) + ΔG f (O 2 )] – [2 ΔG f (H 2 O)]
= 474.4 kJ
ln Kp = –191.5
–191.5
Kp = e
=7 × 10-84 56
Comment This extremely small equilibrium constant is consistent
with the fact that water does not decompose into hydrogen and
oxygen gases at 25°C. Thus a large positive ∆G° favors reactants
over products at equilibrium. Note that we have added the (1
mol) term on the right side of Equation (5-64) here to make the
units consistent.
EXAMPLE 5.18
Using the solubility product of silver chloride at 25°C (1.6 × 10 –10 ),
calculate ∆G° for the process
Solution
Because this is a heterogeneous equilibrium, the solubility product is the
equilibrium constant.
K s p = [Ag + ][CI – ] = 1.6 × 10 –10
The positive value of ΔG° does not imply that the reaction under
consideration may not proceed spontaneously under any conditions. ΔG°
refers to the reaction
where each substance is in its standard state; that is, at a partial pressure
of 1 atm. The positive value of ΔG° only means that the reaction will not
proceed spontaneously under these conditions. However, if we were to start
with isopropyl alcohol at a partial pressure of 1 atm and no acetone or
hydrogen, the alcohol would decompose spontaneously at 452.2 K and more
than 50% dissociation could occur. Yields can be made even better is one of
the products is removed continuously (Le Chatelier’s principle)
We also might calculate ΔG for one set of conditions with the
substances not all in their standard states; for example, 57
(CH3)2CHOH(g, P = 1 atm) ⇄(CH3)2CO(g, P = 0.100 atm) + H2(g, P = 0.100 atm)
For this calculation we refer to Equation (5-63), which relates ΔG to ΔG° and
the non-equilibrium partial pressures
∆G = ∆G° + RT l n Q
Applied to our reaction, we get
P(CH3 ) CO PH2
2
= ΔG + RT ln
P(CH3 ) CHOH
2
(0.100 )2
= 3053 J + (1 mol)(8.314 J/K∙mol)(452.2 K) × ln
1.00
= −15.5 × 103 J
EXAMPLE 5.19
The standard free-energy change for the reaction
N2(g) + 3 H2 ( g) ⇄ 2 NH3(g)
Solution
Equation (7-73) can be written as
∆G = ∆G° + RT ln Qp
2
PNH
3
= ΔG + RT ln
PH3 PN
2 2
(12.9)2
= – 33.2 × 1000 J + (1 mol)(8.314 J / K ⋅ mol )(298 K ) × ln
(0.250)3 (0.870)
58
= – 33.2 × 1000 J + (1 mol)(8.314 J/K ⋅ mol)(298 K) × ln1.22 × 10 4
= –33.2 × 103 J + 23.3 × 103 J
ΔG = ΔH TΔS
ΔG = RT ln K
∆H ∆S
ln K = − + (5-65)
RT R
59
Figure 5.9 lnK versus 1/T for (a) exothermic
reaction with ΔH = 40.0 kJ; (b) endothermic
reaction with ΔH = 85 kJ.
∆H ∆S ∆H ∆S
ln K1 = − + , ln K 2 = − +
RT1 R RT2 R
and then subtracting the first of these expressions from the second. The
result is
K ∆H 1 1
ln 2 = − − (5-66)
K1 R T2 T1
EXAMPLE 5.20
For the reaction
NO(g) + 1/2O2(g) ⇄ NO2(g)
ΔG° = −34.9 kJ and ΔH° = −56.5 kJ at 298 K. Calculate the equilibrium
constant at 298 K and ay 598 K.
Solution
ΔGrxn = − RT ln K
ΔGrxn
ln K = −
RT
− ΔGrxn
K = e RT
K = 1.28 × 106
K2 ΔH 1 1
=−
ln
K1 R T − T
2 1
K2 (−5.65 × 10 4 J) 1 1
ln =− −
1.28 × 10 6 8.314 J/K ⋅ mol 598 298
(−5.65 × 10 4 J) 1 1
ln K 2 − ln(1.28 × 10 6 ) = − −
8.314 J/K ⋅ mol 598 298
K2 = 14.4
The equilibrium constant of this exothermic reaction is smaller at
the higher temperatue, which is consistent with Le Chatelier’s principle. 60
COMMON UNITS AND CONVERSION FACTORS FOR
THERMODYNAMICS
1 cal = 4.184 J
1 L∙atm = 101.29 J
1 atm = 760 mm Hg
1 torr = 1.000 mm Hg
1 torr = 133.32 Pa
1J = 1 kg∙m2∙s–2
1 N (Newton) = kg∙m∙s–2
1 Pa = N∙m–2
1 Pa = kg∙m–1∙s–2
1J = kg∙m2∙s–2
R = 1.9872 cal∙K–1∙mol–1
R = 0.082053 L∙K–1∙mol–1
61
PROBLEMS
II. Using the chemical equations given below, calculate the standard enthalpy, ∆H°, of
the following reaction
HCOOH(l) → CO(g) + H2O(l)
III. Predict whether the following reactions take place spontaneously: a) at 298 K and 1
atm, b) at 800 K and 1 atm
SO3(g) + NO(g) → SO2(g) + NO2(g)
∆H° = 9.98 kcal/mol; ∆S° = 0.01369 kcal/mol. Assume ∆H° and ∆S° do not vary
with temperature.
IV. Using the following equations, calculate the standard free energy of formation, ∆G°
f for ozone at 298 K.
3/2 O2(g) → O3(g)
Given:
∆H°f (O3) = 34.0 kcal/mol
S°(O3) = 56.8 cal/mol.K
S°(O2) = 49.0 cal/mol.K
Assuming that ∆H° and ∆S° are independent of temperature, calculate the
equilibrium constant at 298 K and 600 K.
Given:
∆G°f (SO2 gas) = –71.79 kcal/mol
VIII. The enthalpies of the following reactions are for 0°C and 1 atm:
N2(g) + 3H2(g) → 2NH3(g) ∆H1 = –24.4 kcal
IX. Calculate the bond energy C—H in CH4 using the following data:
CH4(g) + 2O2(g) → CO2(g) + 2H2O(l) ∆H° = –890.5 kJ
XII. For which of the following processes would S have a negative value?
a) 2 Fe2O3(s) → 4 Fe(s) + 3 O2(g)
b) Mg2+(aq) + 2 OH–(aq) → Mg(OH)2(s)
c) 3 H2(g) + 3 C2H4(g) → 3 C2H6(g)
XIII. The standard molar entropy of NH3(g) is 192.45 J / m ol ·K at 298 K and its heat
capacity is given by c = a + bT + cT 2 where a = 29.75 J / m ol ·K b = 25.1 × 10 – 3
J / m ol ·K and c = –1.55 × 10 − 5 J / m ol ·K.
d) C al cul at e t he chan ge i n t he st and ard m ol ar ent rop y, S, from
25ºC at 500ºC.
e) Calculate the standard molar entropy of NH3(g) at 500ºC.
XIV. Make use of the following data on methanol, CH3OH, and a) calculate the entropy
change when 1 mole of liquid methanol at 25°C is heated to its normal boiling point
and then completely evaporated at that temperature. b) Calculate the entropy of 1
mole of gaseous methanol at its boiling point.
Given:
S° for liquid methanol at 25°C = 126.8 J/K·mol
Boiling point of methanol = 64.6°C
ΔHvap =35.25 J/mol
Cp (J/K·mol) = 15.15 + 0.10 × 10−3T − 29.50 × 10−6T2
XVI. Using a flow calorimeter, a chemist studied the hydrogenation of methyl chloride
CH3Cl(g) + H2(g) → CH4(g) + HCl(g)
= –82.32 kJ/mol. Make the adjustment to
and found that at 248°C (541 K), ΔH 521
the standard temperature (298 k) so that this result could be incorporated into tables
of standard data. Cp values at 400 K for this calculation are as follows:
Cp(CH4(g)) = 41,28 J/mol K,
Cp(HCl(g)) = 29.11 J/mol K,
Cp(CH3Cl(g)) = 48.46 J/mol K,
Cp(H2(g)) = 28.56 J/mol K,
XVII. Given the following ∆H°rxn values, expressed as per mol of reactant:
NH3(g) → N2(g) + H2(g) ∆H°1 = + 46.19 kJ ⋅ mol–1
N2O4(g) → 2NO2(g) ∆H°2 = + 58.04 kJ ⋅ mol–1
2 NH3(g) → N2H4(l) + H2(g) ∆H°3 = + 142.8 kJ ⋅ mol–1
2NO2(g) + 2H2(g) → N2(g) + 2H2O2(g) ∆H°4 = – 551.36 kJ ⋅ mol–
Calculate ∆H°rxn at 298 K for the following reaction
N2O4(g) + N2H4(l) → 2N2(g) + 2H2O2(l)
XVIII. Given the thermodynamics data listed in the table below for the following reaction:
4 HCl(g) + O2(g) → 2 H2O(g) + 2 Cl2(g)
, and the
a. Calculate at T1 = 298 K the standard enthalpy change, ΔH rxn
.
change in the standard entropy, ΔS rxn
b. Calculate at T1 = 298 K the internal energy change, ΔErxn . R = 8.314 J/ K
· mol
c. Calculate at T1 = 298 K the standard free energy change, ΔGrxn , and
comment on the significance of the value obtained. Calculate the
equilibrium constant (Kp) for the reaction at 298 K.
65
d. Determine the value of ΔH rxn at T2 = 500 K. Comment briefly on the
significance of the value obtained in relation to the value at T1 (i.e. if it is
consistent with any principal in chemistry.
e. Determine the value of ΔS rxn at T2 = 500 K. Comment briefly on the
significance of the value obtained in relation to the value at T1.
XIX. Using data given below, calculate the standard enthalpy of formation of gaseous
acetaldehyde, CH3COH.
Bond Dissociation Energies of Diatomic Molecules and Average Bond Energies for
Bonds in Polyatomic Molecules in kJ/mol
C—H C—C C=O H—H O=O
+414 +347 +745 +436.4 +498.7
66