You are on page 1of 42

Learning Objectives

 To know the relationship between free energy and the 


equilibrium
 constant.

We have identified three criteria for whether a given reaction will occur spontaneously:
ΔSuniv > 0, ΔGsys < 0, and the relative magnitude of the reaction quotient Q versus the 

equilibrium
 constant K. Recall that if Q < K, then the reaction proceeds spontaneously to the right as written,
resulting in the net conversion of reactants to products. Conversely, if Q > K, then the reaction
proceeds spontaneously to the left as written, resulting in the net conversion of products to
reactants. If Q = K, then the system is at 
equilibrium
, and no net reaction occurs. Table 19.7.119.7.1 summarizes these criteria and their relative
values for spontaneous, nonspontaneous, and 
equilibrium
 processes.
Table 19.7.119.7.1: Criteria for the Spontaneity of a Process as Written
us Equilibrium Nonspontaneo
ection.
ΔSuniv = 0 ΔSuniv < 0
ΔGsys = 0 ΔGsys > 0
Q=K Q>K

Because all three criteria are assessing the same thing—the spontaneity of the process
—it would be most surprising indeed if they were not related. In this section, we
explore the relationship between the standard free energy of reaction (ΔG°) and the 

equilibrium
 constant (K).
Free Energy and the 
Equilibrium
 Constant
Because ΔH° and ΔS° determine the magnitude of ΔG° and because K is a measure of
the ratio of the concentrations of products to the concentrations of reactants, we should
be able to express K in terms of ΔG° and vice versa. "Free Energy", ΔG is equal to the
maximum amount of work a system can perform on its 
surroundings
 while undergoing a spontaneous change. For a reversible process that does not involve
external work, we can express the change in free energy in terms of 
volume
, pressure, entropy, and temperature, thereby eliminating ΔH from the equation for ΔG.
The general relationship can be shown as follow (derivation not shown):
ΔG=VΔP−SΔT(19.7.1)(19.7.1)ΔG=VΔP−SΔT

If a reaction is carried out at constant temperature (ΔT = 0), then


Equation 19.7.119.7.1 simplifies to

ΔG=VΔP(19.7.2)(19.7.2)ΔG=VΔP

Under normal conditions, the pressure dependence of free energy is not important for
solids and liquids because of their small molar volumes. For reactions that involve
gases, however, the effect of pressure on free energy is very important.

Assuming 

ideal gas
 behavior, we can replace the VV in Equation 19.7.219.7.2 by nRT/P (where n is the
number of moles of gas and 
R
 is the 
ideal gas
 constant) and express ΔGΔG in terms of the initial and final pressures (PiPi and PfPf,
respectively):
ΔG=(n
R
TP)ΔP=n
R
TΔPP=n
R
Tln(PfPi)(19.7.3)(19.7.3)ΔG=(n
R
TP)ΔP=n
R
TΔPP=n
R
Tln⁡(PfPi)

If the initial state is the standard state with Pi = 1 atm, then the change in free energy
of a substance when going from the standard state to any other state with a pressure P
can be written as follows:

G−G∘=n
R
TlnP(19.7.4)(19.7.4)G−G°=n
R
Tln⁡P

This can be rearranged as follows:

G=G∘+n
R
TlnP(19.7.5)(19.7.5)G=G°+n
R
Tln⁡P

As you will soon discover, Equation 19.7.519.7.5 allows us to relate ΔG° and Kp. Any
relationship that is true for KpKp must also be true for KK because KpKp and KK are
simply different ways of expressing the 

equilibrium
 constant using different units.

Let’s consider the following hypothetical reaction, in which all the reactants and the
products are ideal gases and the lowercase letters correspond to the stoichiometric
coefficients for the various species:

aA+bB⇌cC+dD(19.7.6)(19.7.6)aA+bB⇌cC+dD

Because the free-energy change for a reaction is the difference between the sum of the
free energies of the products and the reactants, we can write the following expression
for ΔG:

ΔG=∑mGp
r

oducts−∑nG

r
eactants=(cGC+dGD)−(aGA+bGB)(19.7.7)(19.7.7)ΔG=∑mGp
r
oducts−∑nG
r
eactants=(cGC+dGD)−(aGA+bGB)

Substituting Equation 19.7.519.7.5 for each term into Equation 19.7.719.7.7,

ΔG=[(cGoC+c
R
TlnPC)+(dGoD+d
R
TlnPD)]−[(aGoA+a
R
TlnPA)+(bGoB+b
R
TlnPB)](19.7.8)(19.7.8)ΔG=[(cGCo+c
R
Tln⁡PC)+(dGDo+d
R
Tln⁡PD)]−[(aGAo+a
R
Tln⁡PA)+(bGBo+b
R
Tln⁡PB)]

Combining terms gives the following relationship between ΔG and the reaction quotient
Q:

ΔG=ΔG∘+
R
Tln(PcCPdDPaAPbB)=ΔG∘+

R
TlnQ(19.7.9)(19.7.9)ΔG=ΔG∘+
R
Tln⁡(PCcPDdPAaPBb)=ΔG∘+
R
Tln⁡Q
where ΔG° indicates that all reactants and products are in their standard states. For
gases at 

equilibrium
 (Q=KpQ=Kp,), and as you’ve learned in this chapter, ΔG = 0 for a system at 
equilibrium
. Therefore, we can describe the relationship between ΔG° and Kp for gases as follows:
0=ΔG°+
R
TlnKp(19.7.10)(19.7.10)0=ΔG°+
R
Tln⁡Kp
ΔG°=−
R
TlnKp(19.7.11)(19.7.11)ΔG°=−
R
Tln⁡Kp

If the products and reactants are in their standard states and ΔG° < 0, then K p > 1, and
products are favored over reactants when the reaction is at 

equilibrium
. Conversely, if ΔG° > 0, then Kp < 1, and reactants are favored over products when the
reaction is at 
equilibrium
. If ΔG° = 0, then Kp=1Kp=1, and neither reactants nor products are favored when
the reaction is at 
equilibrium
.

For a spontaneous process under standard conditions,  KeqKeq  and  KpKp  are greater
than 1.

Example 19.7.119.7.1

ΔG° is −32.7 kJ/mol of N2 for the reaction

N2(g)+3H2(g)⇌2NH3(g)(19.7.12)(19.7.12)N2(g)+3H2(g)⇌2NH3(g)

This calculation was for the reaction under standard conditions—that is, with all gases
present at a 
partial pressure
 of 1 atm and a temperature of 25°C. Calculate ΔG for the same reaction under the
following nonstandard conditions:
 PN2PN2 = 2.00 atm,
 PH2PH2 = 7.00 atm,
 PNH3PNH3 = 0.021 atm,
 and T = 100°C.

Does the reaction proceed to the right, as written, or to the left to reach 

equilibrium
?

Given: balanced chemical equation, 

partial pressure
 of each species, temperature, and ΔG°

Asked for: whether the reaction proceeds to the right or to the left to reach 

equilibrium

Strategy:

A. Using the values given and Equation 19.7.919.7.9, calculate Q.


B. Determine if Q is >, <, or = to K
C. Substitute the values of ΔG° and Q into Equation 19.7.919.7.9 to obtain ΔG for
the reaction under nonstandard conditions.

Solution:

A The relationship between ΔG° and ΔG under nonstandard conditions is given in


Equation 19.7.919.7.9. Substituting the partial pressures given, we can calculate Q:

Q=P2NH3PN2P3H2=(0.021)2(2.00)(7.00)3=6.4×10−7(19.7.13)
(19.7.13)Q=PNH32PN2PH23=(0.021)2(2.00)(7.00)3=6.4×10−7

B Because ΔG° is −, K must be a number greater than 1

C Substituting the values of ΔG° and Q into Equation 19.7.919.7.9,

ΔG=ΔG∘+
R
TlnQ=−32.7 kJ+[(8.314 J/K)(373 K)(1 kJ1000 J)ln(6.4×10−7)]ΔG=ΔG∘+
R
Tln⁡Q=−32.7 kJ+[(8.314 J/K)(373 K)(1 kJ1000 J)ln⁡(6.4×10−7)]
=−32.7 kJ+(−44 kJ)=−32.7 kJ+(−44 kJ)
=−77 kJ/mol of N2=−77 kJ/mol of N2

Because ΔG < 0 and Q < K (because Q < 1), the reaction proceeds spontaneously to
the right, as written, in order to reach 

equilibrium
.
Exercise 19.7.119.7.1

Calculate ΔG for the reaction of nitric oxide with oxygen to give nitrogen dioxide under
these conditions: T = 50°C, PNO = 0.0100 atm, PO2PO2 = 0.200 atm, and PNO2PNO2 =
1.00 × 10−4 atm. The value of ΔG° for this reaction is −72.5 kJ/mol of O2. Are products
or reactants favored?

Answer
Example 19.7.219.7.2

Calculate Kp for the reaction of H2 with N2 to give NH3 at 25°C. ΔG° for this reaction is
−32.7 kJ/mol of N2.

Given: balanced chemical equation from Example 10, ΔG°, and temperature

Asked for: Kp

Strategy:

Substitute values for ΔG° and T (in kelvin) into Equation 19.7.1119.7.11 to calculate
Kp, the 

equilibrium
 constant for the formation of ammonia.

Solution

In Example 10, we used tabulated values of ΔG∘f to calculate ΔG° for this reaction
(−32.7 kJ/mol of N2). For 

equilibrium
 conditions, rearranging Equation 19.7.1119.7.11,
ΔG∘−ΔG∘
R
T=−
R
TlnKp=lnKp(19.7.14)(19.7.15)(19.7.14)ΔG∘=−
R
Tln⁡Kp(19.7.15)−ΔG∘
R
T=ln⁡Kp

Inserting the value of ΔG° and the temperature (25°C = 298 K) into this equation,

lnKpKp=−(−32.7 kJ)(1000 J/kJ)(8.314 J/K)(298 K)=13.2=5.4×105(19.7.16)


(19.7.17)(19.7.16)ln⁡Kp=−(−32.7 kJ)(1000 J/kJ)(8.314 J/K)(298
K)=13.2(19.7.17)Kp=5.4×105

Thus the 

equilibrium
 constant for the formation of ammonia at room temperature is product-favored.
However, the rate at which the reaction occurs at room temperature is too slow to be
useful.
Exercise 19.7.319.7.3

Calculate Kp for the reaction of NO with O2 to give NO2 at 25°C. ΔG° for this reaction is
−70.5 kJ/mol of O2.

Answer

Although Kp is defined in terms of the partial pressures of the reactants and the
products, the 

equilibrium
 constant K is defined in terms of the concentrations of the reactants and the products.
We described the relationship between the numerical magnitude of Kp and K in Chapter
15 and showed that they are related:
Kp=K(
R
T)Δn(19.7.18)(19.7.18)Kp=K(
R
T)Δn
where Δn is the number of moles of gaseous product minus the number of moles of
gaseous reactant. For reactions that involve only 

solutions
, liquids, and solids, Δn = 0, so Kp = K. For all reactions that do not involve a change in
the number of moles of gas present, the relationship in Equation 19.7.1119.7.11 can
be written in a more general form:
ΔG°=−
R
TlnK(19.7.19)(19.7.19)ΔG°=−
R
Tln⁡K

Only when a reaction results in a net production or consumption of gases is it necessary


to correct Equation 19.7.1919.7.19 for the difference between Kp and K.

Non-Ideal Behavior

Although we typically use concentrations or pressures in our 

equilibrium
 calculations, recall that 
equilibrium
 constants are generally expressed as unitless numbers because of the use
of activities or fugacities in 
precise
 thermodynamic work. Systems that contain gases at high pressures or concentrated 
solutions
 that deviate substantially from ideal behavior require the use of fugacities or activities,
respectively.

Combining
Equation 19.7.1919.7.19 with ΔGo=ΔHo−TΔSoΔGo=ΔHo−TΔSo provides insight
into how the components of ΔG° influence the magnitude of the 

equilibrium
 constant:
ΔG°=ΔH°−TΔS°=−
R
TlnK(19.7.20)(19.7.20)ΔG°=ΔH°−TΔS°=−
R
Tln⁡K
Notice that KK becomes larger as ΔS° becomes more positive, indicating that the
magnitude of the 

equilibrium
 constant is directly influenced by the tendency of a system to move toward maximum
disorder. Moreover, K increases as ΔH° decreases. Thus the magnitude of the 
equilibrium
 constant is also directly influenced by the tendency of a system to seek the lowest
energy state possible.

The magnitude of the 

equilibrium
  constant is directly influenced by the tendency of a system to move toward maximum
entropy and seek the lowest energy state possible.

Temperature Dependence of the 


Equilibrium
 Constant
The fact that ΔG° and K are related provides us with another explanation of why 

equilibrium
 constants are temperature dependent. This relationship is shown explicitly in
Equation 19.7.2019.7.20, which can be rearranged as follows:
lnK=−ΔH∘
R
T+ΔS∘
R
(19.7.21)(19.7.21)ln⁡K=−ΔH∘
R
T+ΔS∘
R

Assuming ΔH° and ΔS° are temperature independent, for an 

exothermic
 reaction (ΔH° < 0), the magnitude of K decreases with increasing temperature,
whereas for an 
endothermic
 reaction (ΔH° > 0), the magnitude of K increases with increasing temperature. The
quantitative relationship expressed in Equation 19.7.2119.7.21 agrees with the
qualitative predictions made by applying Le Chatelier’s principle. Because heat is
produced in an 
exothermic
 reaction, adding heat (by increasing the temperature) will shift the 
equilibrium
 to the left, favoring the reactants and decreasing the magnitude of K. Conversely,
because heat is consumed in an 
endothermic
 reaction, adding heat will shift the 
equilibrium
 to the right, favoring the products and increasing the magnitude of K.
Equation 19.7.2119.7.21 also shows that the magnitude of ΔH° dictates how rapidly K
changes as a function of temperature. In contrast, the magnitude and sign of ΔS° affect
the magnitude of K but not its temperature dependence.

If we know the value of K at a given temperature and the value of ΔH° for a reaction,
we can estimate the value of K at any other temperature, even in the absence of
information on ΔS°. Suppose, for example, that K1 and K2 are the 

equilibrium
 constants for a reaction at temperatures T1 and T2, respectively. Applying
Equation 19.7.2119.7.21 gives the following relationship at each temperature:
lnK1lnK2=−ΔH∘
R
T1+ΔS∘
R
=−ΔH∘
R
T2+ΔS∘
R
(19.7.22)(19.7.23)(19.7.22)ln⁡K1=−ΔH∘
R
T1+ΔS∘
R
(19.7.23)ln⁡K2=−ΔH∘
R
T2+ΔS∘
R

Subtracting lnK1ln⁡K1 from lnK2ln⁡K2,
lnK2−lnK1=lnK2K1=ΔH∘
R
(1T1−1T2)(19.7.24)(19.7.24)ln⁡K2−ln⁡K1=ln⁡K2K1=ΔH∘
R
(1T1−1T2)

Thus calculating ΔH° from tabulated enthalpies of formation and measuring the 

equilibrium
 constant at one temperature (K1) allow us to calculate the value of the 
equilibrium
 constant at any other temperature (K2), assuming that ΔH° and ΔS° are independent of
temperature.
Example 19.7.419.7.4

The 

equilibrium
 constant for the formation of NH3 from H2 and N2 at 25°C was calculated to be Kp = 5.4
× 105 in Example 13. What is Kp at 500°C? (Use the data from Example 10.)

Given: balanced chemical equation, ΔH°, initial and final T, and Kp at 25°C

Asked for: Kp at 500°C

Strategy:

Convert the initial and final temperatures to kelvins. Then substitute appropriate values
into Equation 19.7.2419.7.24 to obtain K2, the 

equilibrium
 constant at the final temperature.

Solution:

The value of ΔH° for the reaction obtained using Hess’s law is −91.8 kJ/mol of N 2. If we
set T1 = 25°C = 298.K and T2 = 500°C = 773 K, then from
Equation 19.7.2419.7.24 we obtain the following:

lnK2K1K2K1K2=ΔH∘
R
(1T1−1T2)=(−91.8 kJ)(1000 J/kJ)8.314 J/K(1298 K−1773

K)=−22.8=1.3×10−10=(5.4×105)(1.3×10−10)=7.0×10−5(19.7.25)(19.7.26)
(19.7.27)(19.7.28)(19.7.25)ln⁡K2K1=ΔH∘
R
(1T1−1T2)(19.7.26)=(−91.8 kJ)(1000 J/kJ)8.314 J/K(1298 K−1773
K)=−22.8(19.7.27)K2K1=1.3×10−10(19.7.28)K2=(5.4×105)(1.3×10−10)=7.0×10−5

Thus at 500°C, the 

equilibrium
 strongly favors the reactants over the products.
Exercise 19.7.419.7.4

In the exercise in Example 19.7.319.7.3, you calculated Kp = 2.2 × 1012 for the reaction
of NO with O2 to give NO2 at 25°C. Use the ΔH∘f values in the exercise in Example 10 to
calculate Kp for this reaction at 1000°C.

Answer

Summary
For a reversible process that does not involve external work, we can express the
change in free energy in terms of 

volume
, pressure, entropy, and temperature. If we assume 
ideal gas
 behavior, the 
ideal gas law
 allows us to express ΔG in terms of the partial pressures of the reactants and products,
which gives us a relationship between ΔG and Kp, the 
equilibrium
 constant of a reaction involving gases, or K, the 
equilibrium
 constant expressed in terms of concentrations. If ΔG° < 0, then K > 1, and products
are favored over reactants at 
equilibrium
. Conversely, if ΔG° > 0, then K < 1, and reactants are favored over products at 
equilibrium
. If ΔG° = 0, then K=1, and neither reactants nor products are favored at 
equilibrium
. We can use the measured 
equilibrium
 constant K at one temperature and ΔH° to estimate the 
equilibrium
 constant for a reaction at any other temperature.

Learning Objectives
 To know the relationship between free energy and the 
equilibrium
 constant.

We have identified three criteria for whether a given reaction will occur spontaneously:
ΔSuniv > 0, ΔGsys < 0, and the relative magnitude of the reaction quotient Q versus the 

equilibrium
 constant K. Recall that if Q < K, then the reaction proceeds spontaneously to the right as written,
resulting in the net conversion of reactants to products. Conversely, if Q > K, then the reaction
proceeds spontaneously to the left as written, resulting in the net conversion of products to
reactants. If Q = K, then the system is at 
equilibrium
, and no net reaction occurs. Table 19.7.119.7.1 summarizes these criteria and their relative
values for spontaneous, nonspontaneous, and 
equilibrium
 processes.
Table 19.7.119.7.1: Criteria for the Spontaneity of a Process as Written
us Equilibrium Nonspontaneo
ection.
ΔSuniv = 0 ΔSuniv < 0
ΔGsys = 0 ΔGsys > 0
Q=K Q>K

Because all three criteria are assessing the same thing—the spontaneity of the process
—it would be most surprising indeed if they were not related. In this section, we
explore the relationship between the standard free energy of reaction (ΔG°) and the 

equilibrium
 constant (K).
Free Energy and the 
Equilibrium
 Constant
Because ΔH° and ΔS° determine the magnitude of ΔG° and because K is a measure of
the ratio of the concentrations of products to the concentrations of reactants, we should
be able to express K in terms of ΔG° and vice versa. "Free Energy", ΔG is equal to the
maximum amount of work a system can perform on its 

surroundings
 while undergoing a spontaneous change. For a reversible process that does not involve
external work, we can express the change in free energy in terms of 
volume
, pressure, entropy, and temperature, thereby eliminating ΔH from the equation for ΔG.
The general relationship can be shown as follow (derivation not shown):
ΔG=VΔP−SΔT(19.7.1)(19.7.1)ΔG=VΔP−SΔT

If a reaction is carried out at constant temperature (ΔT = 0), then


Equation 19.7.119.7.1 simplifies to

ΔG=VΔP(19.7.2)(19.7.2)ΔG=VΔP

Under normal conditions, the pressure dependence of free energy is not important for
solids and liquids because of their small molar volumes. For reactions that involve
gases, however, the effect of pressure on free energy is very important.

Assuming 

ideal gas
 behavior, we can replace the VV in Equation 19.7.219.7.2 by nRT/P (where n is the
number of moles of gas and 
R
 is the 
ideal gas
 constant) and express ΔGΔG in terms of the initial and final pressures (PiPi and PfPf,
respectively):
ΔG=(n
R
TP)ΔP=n
R
TΔPP=n
R
Tln(PfPi)(19.7.3)(19.7.3)ΔG=(n
R
TP)ΔP=n
R
TΔPP=n
R
Tln⁡(PfPi)

If the initial state is the standard state with Pi = 1 atm, then the change in free energy
of a substance when going from the standard state to any other state with a pressure P
can be written as follows:

G−G∘=n
R
TlnP(19.7.4)(19.7.4)G−G°=n
R
Tln⁡P

This can be rearranged as follows:

G=G∘+n
R
TlnP(19.7.5)(19.7.5)G=G°+n
R
Tln⁡P

As you will soon discover, Equation 19.7.519.7.5 allows us to relate ΔG° and Kp. Any
relationship that is true for KpKp must also be true for KK because KpKp and KK are
simply different ways of expressing the 

equilibrium
 constant using different units.

Let’s consider the following hypothetical reaction, in which all the reactants and the
products are ideal gases and the lowercase letters correspond to the stoichiometric
coefficients for the various species:

aA+bB⇌cC+dD(19.7.6)(19.7.6)aA+bB⇌cC+dD
Because the free-energy change for a reaction is the difference between the sum of the
free energies of the products and the reactants, we can write the following expression
for ΔG:

ΔG=∑mGp
r

oducts−∑nG

r
eactants=(cGC+dGD)−(aGA+bGB)(19.7.7)(19.7.7)ΔG=∑mGp
r
oducts−∑nG
r
eactants=(cGC+dGD)−(aGA+bGB)

Substituting Equation 19.7.519.7.5 for each term into Equation 19.7.719.7.7,

ΔG=[(cGoC+c
R
TlnPC)+(dGoD+d
R
TlnPD)]−[(aGoA+a
R
TlnPA)+(bGoB+b
R
TlnPB)](19.7.8)(19.7.8)ΔG=[(cGCo+c
R
Tln⁡PC)+(dGDo+d
R
Tln⁡PD)]−[(aGAo+a
R
Tln⁡PA)+(bGBo+b
R
Tln⁡PB)]

Combining terms gives the following relationship between ΔG and the reaction quotient
Q:

ΔG=ΔG∘+
R
Tln(PcCPdDPaAPbB)=ΔG∘+

R
TlnQ(19.7.9)(19.7.9)ΔG=ΔG∘+
R
Tln⁡(PCcPDdPAaPBb)=ΔG∘+
R
Tln⁡Q

where ΔG° indicates that all reactants and products are in their standard states. For
gases at 

equilibrium
 (Q=KpQ=Kp,), and as you’ve learned in this chapter, ΔG = 0 for a system at 
equilibrium
. Therefore, we can describe the relationship between ΔG° and Kp for gases as follows:
0=ΔG°+
R
TlnKp(19.7.10)(19.7.10)0=ΔG°+
R
Tln⁡Kp
ΔG°=−
R
TlnKp(19.7.11)(19.7.11)ΔG°=−
R
Tln⁡Kp

If the products and reactants are in their standard states and ΔG° < 0, then K p > 1, and
products are favored over reactants when the reaction is at 

equilibrium
. Conversely, if ΔG° > 0, then Kp < 1, and reactants are favored over products when the
reaction is at 
equilibrium
. If ΔG° = 0, then Kp=1Kp=1, and neither reactants nor products are favored when
the reaction is at 
equilibrium
.

For a spontaneous process under standard conditions,  KeqKeq  and  KpKp  are greater
than 1.
Example 19.7.119.7.1

ΔG° is −32.7 kJ/mol of N2 for the reaction

N2(g)+3H2(g)⇌2NH3(g)(19.7.12)(19.7.12)N2(g)+3H2(g)⇌2NH3(g)

This calculation was for the reaction under standard conditions—that is, with all gases
present at a 

partial pressure
 of 1 atm and a temperature of 25°C. Calculate ΔG for the same reaction under the
following nonstandard conditions:
 PN2PN2 = 2.00 atm,
 PH2PH2 = 7.00 atm,
 PNH3PNH3 = 0.021 atm,
 and T = 100°C.

Does the reaction proceed to the right, as written, or to the left to reach 

equilibrium
?

Given: balanced chemical equation, 

partial pressure
 of each species, temperature, and ΔG°

Asked for: whether the reaction proceeds to the right or to the left to reach 

equilibrium

Strategy:

A. Using the values given and Equation 19.7.919.7.9, calculate Q.


B. Determine if Q is >, <, or = to K
C. Substitute the values of ΔG° and Q into Equation 19.7.919.7.9 to obtain ΔG for
the reaction under nonstandard conditions.

Solution:

A The relationship between ΔG° and ΔG under nonstandard conditions is given in


Equation 19.7.919.7.9. Substituting the partial pressures given, we can calculate Q:
Q=P2NH3PN2P3H2=(0.021)2(2.00)(7.00)3=6.4×10−7(19.7.13)
(19.7.13)Q=PNH32PN2PH23=(0.021)2(2.00)(7.00)3=6.4×10−7

B Because ΔG° is −, K must be a number greater than 1

C Substituting the values of ΔG° and Q into Equation 19.7.919.7.9,

ΔG=ΔG∘+
R
TlnQ=−32.7 kJ+[(8.314 J/K)(373 K)(1 kJ1000 J)ln(6.4×10−7)]ΔG=ΔG∘+
R
Tln⁡Q=−32.7 kJ+[(8.314 J/K)(373 K)(1 kJ1000 J)ln⁡(6.4×10−7)]
=−32.7 kJ+(−44 kJ)=−32.7 kJ+(−44 kJ)
=−77 kJ/mol of N2=−77 kJ/mol of N2

Because ΔG < 0 and Q < K (because Q < 1), the reaction proceeds spontaneously to
the right, as written, in order to reach 

equilibrium
.
Exercise 19.7.119.7.1

Calculate ΔG for the reaction of nitric oxide with oxygen to give nitrogen dioxide under
these conditions: T = 50°C, PNO = 0.0100 atm, PO2PO2 = 0.200 atm, and PNO2PNO2 =
1.00 × 10−4 atm. The value of ΔG° for this reaction is −72.5 kJ/mol of O2. Are products
or reactants favored?

Answer
Example 19.7.219.7.2

Calculate Kp for the reaction of H2 with N2 to give NH3 at 25°C. ΔG° for this reaction is
−32.7 kJ/mol of N2.

Given: balanced chemical equation from Example 10, ΔG°, and temperature

Asked for: Kp

Strategy:

Substitute values for ΔG° and T (in kelvin) into Equation 19.7.1119.7.11 to calculate
Kp, the 
equilibrium
 constant for the formation of ammonia.

Solution

In Example 10, we used tabulated values of ΔG∘f to calculate ΔG° for this reaction
(−32.7 kJ/mol of N2). For 

equilibrium
 conditions, rearranging Equation 19.7.1119.7.11,
ΔG∘−ΔG∘
R
T=−
R
TlnKp=lnKp(19.7.14)(19.7.15)(19.7.14)ΔG∘=−
R
Tln⁡Kp(19.7.15)−ΔG∘
R
T=ln⁡Kp

Inserting the value of ΔG° and the temperature (25°C = 298 K) into this equation,

lnKpKp=−(−32.7 kJ)(1000 J/kJ)(8.314 J/K)(298 K)=13.2=5.4×105(19.7.16)


(19.7.17)(19.7.16)ln⁡Kp=−(−32.7 kJ)(1000 J/kJ)(8.314 J/K)(298
K)=13.2(19.7.17)Kp=5.4×105

Thus the 

equilibrium
 constant for the formation of ammonia at room temperature is product-favored.
However, the rate at which the reaction occurs at room temperature is too slow to be
useful.
Exercise 19.7.319.7.3

Calculate Kp for the reaction of NO with O2 to give NO2 at 25°C. ΔG° for this reaction is
−70.5 kJ/mol of O2.

Answer

Although Kp is defined in terms of the partial pressures of the reactants and the
products, the 

equilibrium
 constant K is defined in terms of the concentrations of the reactants and the products.
We described the relationship between the numerical magnitude of Kp and K in Chapter
15 and showed that they are related:
Kp=K(
R
T)Δn(19.7.18)(19.7.18)Kp=K(
R
T)Δn

where Δn is the number of moles of gaseous product minus the number of moles of
gaseous reactant. For reactions that involve only 

solutions
, liquids, and solids, Δn = 0, so Kp = K. For all reactions that do not involve a change in
the number of moles of gas present, the relationship in Equation 19.7.1119.7.11 can
be written in a more general form:
ΔG°=−
R
TlnK(19.7.19)(19.7.19)ΔG°=−
R
Tln⁡K

Only when a reaction results in a net production or consumption of gases is it necessary


to correct Equation 19.7.1919.7.19 for the difference between Kp and K.

Non-Ideal Behavior

Although we typically use concentrations or pressures in our 

equilibrium
 calculations, recall that 
equilibrium
 constants are generally expressed as unitless numbers because of the use
of activities or fugacities in 
precise
 thermodynamic work. Systems that contain gases at high pressures or concentrated 
solutions
 that deviate substantially from ideal behavior require the use of fugacities or activities,
respectively.

Combining
Equation 19.7.1919.7.19 with ΔGo=ΔHo−TΔSoΔGo=ΔHo−TΔSo provides insight
into how the components of ΔG° influence the magnitude of the 
equilibrium
 constant:
ΔG°=ΔH°−TΔS°=−
R
TlnK(19.7.20)(19.7.20)ΔG°=ΔH°−TΔS°=−
R
Tln⁡K

Notice that KK becomes larger as ΔS° becomes more positive, indicating that the
magnitude of the 

equilibrium
 constant is directly influenced by the tendency of a system to move toward maximum
disorder. Moreover, K increases as ΔH° decreases. Thus the magnitude of the 
equilibrium
 constant is also directly influenced by the tendency of a system to seek the lowest
energy state possible.

The magnitude of the 

equilibrium
  constant is directly influenced by the tendency of a system to move toward maximum
entropy and seek the lowest energy state possible.

Temperature Dependence of the 


Equilibrium
 Constant
The fact that ΔG° and K are related provides us with another explanation of why 

equilibrium
 constants are temperature dependent. This relationship is shown explicitly in
Equation 19.7.2019.7.20, which can be rearranged as follows:
lnK=−ΔH∘
R
T+ΔS∘
R
(19.7.21)(19.7.21)ln⁡K=−ΔH∘
R
T+ΔS∘
R

Assuming ΔH° and ΔS° are temperature independent, for an 

exothermic
 reaction (ΔH° < 0), the magnitude of K decreases with increasing temperature,
whereas for an 
endothermic
 reaction (ΔH° > 0), the magnitude of K increases with increasing temperature. The
quantitative relationship expressed in Equation 19.7.2119.7.21 agrees with the
qualitative predictions made by applying Le Chatelier’s principle. Because heat is
produced in an 
exothermic
 reaction, adding heat (by increasing the temperature) will shift the 
equilibrium
 to the left, favoring the reactants and decreasing the magnitude of K. Conversely,
because heat is consumed in an 
endothermic
 reaction, adding heat will shift the 
equilibrium
 to the right, favoring the products and increasing the magnitude of K.
Equation 19.7.2119.7.21 also shows that the magnitude of ΔH° dictates how rapidly K
changes as a function of temperature. In contrast, the magnitude and sign of ΔS° affect
the magnitude of K but not its temperature dependence.

If we know the value of K at a given temperature and the value of ΔH° for a reaction,
we can estimate the value of K at any other temperature, even in the absence of
information on ΔS°. Suppose, for example, that K1 and K2 are the 

equilibrium
 constants for a reaction at temperatures T1 and T2, respectively. Applying
Equation 19.7.2119.7.21 gives the following relationship at each temperature:
lnK1lnK2=−ΔH∘
R
T1+ΔS∘
R
=−ΔH∘
R
T2+ΔS∘
R
(19.7.22)(19.7.23)(19.7.22)ln⁡K1=−ΔH∘
R
T1+ΔS∘
R
(19.7.23)ln⁡K2=−ΔH∘
R
T2+ΔS∘
R

Subtracting lnK1ln⁡K1 from lnK2ln⁡K2,

lnK2−lnK1=lnK2K1=ΔH∘
R
(1T1−1T2)(19.7.24)(19.7.24)ln⁡K2−ln⁡K1=ln⁡K2K1=ΔH∘
R
(1T1−1T2)

Thus calculating ΔH° from tabulated enthalpies of formation and measuring the 

equilibrium
 constant at one temperature (K1) allow us to calculate the value of the 
equilibrium
 constant at any other temperature (K2), assuming that ΔH° and ΔS° are independent of
temperature.
Example 19.7.419.7.4

The 

equilibrium
 constant for the formation of NH3 from H2 and N2 at 25°C was calculated to be Kp = 5.4
× 105 in Example 13. What is Kp at 500°C? (Use the data from Example 10.)

Given: balanced chemical equation, ΔH°, initial and final T, and Kp at 25°C

Asked for: Kp at 500°C

Strategy:

Convert the initial and final temperatures to kelvins. Then substitute appropriate values
into Equation 19.7.2419.7.24 to obtain K2, the 

equilibrium
 constant at the final temperature.
Solution:

The value of ΔH° for the reaction obtained using Hess’s law is −91.8 kJ/mol of N 2. If we
set T1 = 25°C = 298.K and T2 = 500°C = 773 K, then from
Equation 19.7.2419.7.24 we obtain the following:

lnK2K1K2K1K2=ΔH∘
R
(1T1−1T2)=(−91.8 kJ)(1000 J/kJ)8.314 J/K(1298 K−1773

K)=−22.8=1.3×10−10=(5.4×105)(1.3×10−10)=7.0×10−5(19.7.25)(19.7.26)
(19.7.27)(19.7.28)(19.7.25)ln⁡K2K1=ΔH∘
R
(1T1−1T2)(19.7.26)=(−91.8 kJ)(1000 J/kJ)8.314 J/K(1298 K−1773
K)=−22.8(19.7.27)K2K1=1.3×10−10(19.7.28)K2=(5.4×105)(1.3×10−10)=7.0×10−5

Thus at 500°C, the 

equilibrium
 strongly favors the reactants over the products.
Exercise 19.7.419.7.4

In the exercise in Example 19.7.319.7.3, you calculated Kp = 2.2 × 1012 for the reaction
of NO with O2 to give NO2 at 25°C. Use the ΔH∘f values in the exercise in Example 10 to
calculate Kp for this reaction at 1000°C.

Answer

Summary
For a reversible process that does not involve external work, we can express the
change in free energy in terms of 

volume
, pressure, entropy, and temperature. If we assume 
ideal gas
 behavior, the 
ideal gas law
 allows us to express ΔG in terms of the partial pressures of the reactants and products,
which gives us a relationship between ΔG and Kp, the 
equilibrium
 constant of a reaction involving gases, or K, the 
equilibrium
 constant expressed in terms of concentrations. If ΔG° < 0, then K > 1, and products
are favored over reactants at 
equilibrium
. Conversely, if ΔG° > 0, then K < 1, and reactants are favored over products at 
equilibrium
. If ΔG° = 0, then K=1, and neither reactants nor products are favored at 
equilibrium
. We can use the measured 
equilibrium
 constant K at one temperature and ΔH° to estimate the 
equilibrium
 constant for a reaction at any other temperature.

Section 2.4Biochemical Energetics

The production of energy, its storage, and its use are as central to the economy of the cell as they are to
the management of the world’s resources. Cells require energy to do all their work, including the
synthesis of sugars from carbon dioxide and water in photosynthesis, the contraction of muscles, and
the replication of DNA. Energy may be defined as the ability to do work, a concept that is easy to grasp
when it is applied to automobile engines and electric power plants. When we consider the energy
associated with chemical bonds and chemical reactions within cells, however, the concept of work
becomes less intuitive.

Go to:

Living Systems Use Various Forms of Energy, Which Are Interconvertible

There are two principal forms of energy: kinetic and potential. Kinetic energy is the energy of
movement — the motion of molecules, for example. The second form of energy, potential energy, or
stored energy, is more important in the study of biological or chemical systems.

Kinetic Energy

Heat, or thermal energy, is a form of kinetic energy — the energy of the motion of molecules. For heat
to do work, it must flow from a region of higher temperature — where the average speed of molecular
motion is greater — to one of lower temperature. Differences in temperature often exist between the
internal and external environments of cells; however, cells generally cannot harness these heat
differentials to do work. Even in warm-blooded animals that have evolved a mechanism for
thermoregulation, the kinetic energy of molecules is used chiefly to maintain constant organismic
temperatures.

Radiant energy is the kinetic energy of photons, or waves of light, and is critical to biology. Radiant
energy can be converted to thermal energy, for instance when light is absorbed by molecules and the
energy is converted to molecular motion. In the process of photosynthesis, light energy is absorbed by
chlorophyll and is ultimately converted into other types of energy, such as that stored in covalent
chemical bonds.

One of the major forms of electric energy is also kinetic — the energy of moving electrons or other
charged particles.

Potential Energy

Several forms of potential energy are biologically significant. Central to biology is the potential energy
stored in the bonds connecting atoms in molecules. Indeed, most of the biochemical reactions described
in this book involve the making or breaking of at least one covalent chemical bond. We recognize this
energy when chemicals undergo energy-releasing reactions. The sugar glucose, for example, is high in
potential energy. Cells degrade glucose continuously, and the energy released when glucose is
metabolized is harnessed to do many kinds of work.

A second biologically important form of potential energy, to which we shall refer often, is the energy in a
concentration gradient. When the concentration of a substance on one side of a permeable barrier, such
as a membrane, is different from that on the other side, the result is a concentration gradient. All cells
form concentration gradients between their interior and the external fluids by selectively exchanging
nutrients, waste products, and ions with their surroundings. Also, compartments within cells frequently
contain different concentrations of ions and other molecules; the concentration of protons within a
lysosome, as we saw in the last section, is about 500 times that of the cytosol.

A third form of potential energy in cells is an electric potential — the energy of charge separation. For
instance, there is a gradient of electric charge of ≈200,000 volts per cm across the outer, or “plasma,”
membrane of virtually all cells.

Interconvertibility of All Forms of Energy

According to the first law of thermodynamics, energy is neither created nor destroyed, but can be
converted from one form to another.* In photosynthesis, for example, as we have just seen, the radiant
energy of light is transformed into the chemical potential energy of the covalent bonds between the
atoms in a sucrose or starch molecule. In muscles and nerves, chemical potential energy stored in
covalent bonds is transformed, respectively, into kinetic and electric energy. In all cells, chemical
potential energy, released by breakage of certain chemical bonds, is used to generate potential energy
in the form of concentration and electric potential gradients. Similarly, energy stored in chemical
concentration gradients or electric potential gradients is used to synthesize chemical bonds, or to
transport other molecules “uphill” against a concentration gradient. This latter process occurs during the
transport of nutrients such as glucose into certain cells and transport of many waste products out of
cells. Because all forms of energy are interconvertible, they can be expressed in the same units of
measurement, such as the calorie or kilocalorie.

Go to:

The Change in Free Energy ΔG Determines the Direction of a Chemical Reaction

Because biological systems are generally held at constant temperature and pressure, it is possible to
predict the direction of a chemical reaction by using a measure of potential energy called free energy, or
G, after the great American chemist Josiah Willard Gibbs (1839 – 1903), a founder of the science of
thermodynamics. Gibbs showed that under conditions of constant pressure and temperature, as
generally found in biological systems, “all systems change in such a way that free energy is minimized.”
In general, we are interested in what happens to the free energy when one molecule or molecular
configuration is changed into another. Thus our concern is with relative, rather than absolute, values of
free energy — in particular, with the difference between the values before and after the change. This
free-energy change ΔG, where Δ stands for difference, is given by

Image ch2e53.jpg

In mathematical terms, Gibbs’s law — that systems change to minimize free energy — is a set of
statements about ΔG:

If ΔG is negative for a chemical reaction or mechanical process, the forward reaction or process (from
left to right as written) will tend to occur spontaneously.

If ΔG is positive, the reverse reaction (from right to left as written) will tend to occur.

If ΔG is zero, both forward and reverse reactions occur at equal rates; the reaction is at equilibrium.

The value of ΔG, like the equilibrium constant, is independent of the reaction mechanism and rate.
Reactions with negative ΔG values that have very slow rate constants may not occur, for practical
purposes, unless a catalyst is present, but the presence of a catalyst does not affect the value of ΔG.

Go to:

The ΔG of a Reaction Depends on Changes in Enthalpy (Bond Energy) and Entropy


At any constant temperature and pressure, two factors determine the ΔG of a reaction and thus
whether the reaction will tend to occur: the change in bond energy between reactants and products and
the change in the randomness of the system. Gibbs showed that free energy can be defined as

Image ch2e54.jpg

where H is the bond energy, or enthalpy, of the system; T is its temperature in degrees Kelvin (K); and S
is a measure of randomness, called entropy. If temperature remains constant, a reaction proceeds
spontaneously only if the freeenergy change ΔG in the following equation is negative:

Image ch2e55.jpg

The enthalpy H of reactants or of products is equal to their total bond energies; the overall change in
enthalpy ΔH is equal to the overall change in bond energies (see Table 2-1). In an exothermic reaction,
the products contain less bond energy than the reactants, the liberated energy is usually converted to
heat (the energy of molecular motion), and ΔH is negative. In an endothermic reaction, the products
contain more bond energy than the reactants, heat is absorbed, and ΔH is positive. Reactions tend to
proceed if they liberate energy (if ΔH < 0), but this is only one of two important parameters of free
energy to consider; the other is entropy.

Entropy S is a measure of the degree of randomness or disorder of a system. Entropy increases as a


system becomes more disordered and decreases as it becomes more structured. Consider, for example,
the diffusion of solutes from one solution into another one in which their concentration is lower. This
important biological reaction is driven only by an increase in entropy; in such a process ΔH is near zero.
To see this, suppose that a 0.1 M solution of glucose is separated from a large volume of water by a
membrane through which glucose can diffuse. Diffusion of glucose molecules across the membrane will
give them more room in which to move, with the result that the randomness, or entropy, of the system
is increased. Maximum entropy is achieved when all molecules can diffuse freely over the largest
possible volume — that is, when the concentration of glucose molecules is the same on both sides of
the membrane. If the degree of hydration of glucose does not change significantly on dilution, ΔH will be
approximately zero; the negative free energy of the reaction in which glucose molecules are liberated to
diffuse over a larger volume will be due solely to the positive value of ΔS in Equation 2-7.

As mentioned previously, the formation of hydrophobic bonds is driven primarily by a change in


entropy. That is, if a long hydrophobic molecule, such as heptane or tristearin, is dissolved in water, the
water molecules are forced to form a cage around it, restricting their free motion. This imposes a high
degree of order on their arrangement and lowers the entropy of the system (ΔS < 0). Because the
entropy change is negative, hydrophobic molecules do not dissolve well in aqueous solutions and tend
to stay associated with one another.

We can summarize the relationships between free energy, enthalpy, and entropy as follows:
An exothermic reaction (ΔH < 0) that increases entropy (ΔS > 0) occurs spontaneously (ΔG < 0).

An endothermic reaction (ΔH > 0) will occur spontaneously if ΔS increases enough so that the T ΔS term
can overcome the positive ΔH.

If the conversion of reactants into products results in no change in free energy (ΔG = 0), then the system
is at equilibrium; that is, any conversion of reactants to products is balanced by an equal conversion of
products to reactants.

Many biological reactions lead to an increase in order, and thus a decrease in entropy (ΔS < 0). An
obvious example is the reaction that links amino acids together to form a protein. A solution of protein
molecules has a lower entropy than does a solution of the same amino acids unlinked, because the free
movement of any amino acid in a protein is restricted when it is bound in a long chain. For the linking
reaction to proceed, a compensatory decrease in free energy must occur elsewhere in the system, as is
discussed in Chapter 4.

Go to:

Several Parameters Affect the ΔG of a Reaction

The change in free energy of a reaction (ΔG) is influenced by temperature, pressure, and the initial
concentrations of reactants and products. Most biological reactions — like others that take place in
aqueous solutions — also are affected by the pH of the solution.

The standard free-energy change of a reaction ΔG°′ is the value of the change in free energy under the
conditions of 298 K (25 °C), 1 atm pressure, pH 7.0 (as in pure water), and initial concentrations of 1 M
for all reactants and products except protons, which are kept at pH 7.0. Table 2-4 gives values of ΔG°′ for
some typical biochemical reactions. The sign of ΔG°′ depends on the direction in which the reaction is
written. If the reaction A → B has a ΔG°′ of −x kcal/mol, then the reverse reaction B → A will have a ΔG°′
value of +x kcal/mol.

Table 2-4. Values of ΔG°′, the Standard Free-Energy Change, for Some Important Biochemical Reactions.

Table 2-4

Values of ΔG°′, the Standard Free-Energy Change, for Some Important Biochemical Reactions.

Most biological reactions differ from standard conditions, particularly in the concentrations of reactants.
However, we can estimate free-energy changes for different temperatures and initial concentrations,
using the equation

Image ch2e56.jpg
where R is the gas constant of 1.987 cal/(degree · mol), T is the temperature (in degrees Kelvin), and Q is
the initial ratio of products to reactants, which is expressed as in Equation 2-1 defining the equilibrium
constant. Again using as our example the interconversion of glyceraldehyde 3-phosphate (G3P) and
dihydroxyacetone phosphate (DHAP)

Image ch2e57.jpg

we have Q = [DHAP]/[G3P] and ΔG°′ = −1840 cal/mol (see Table 2-4). Equation 2-8 for ΔG then becomes

Image ch2e58.jpg

from which we can calculate ΔG for any set of concentrations of DHAP and G3P. If the initial
concentrations of both DHAP and G3P are 1 M, then ΔG = ΔG°′ =  −1840 cal/mol, because RT ln 1 = 0.
The reaction will tend to proceed from left to right, in the direction of formation of DHAP. If, however,
the initial concentration of DHAP is 0.1 M and that of G3P is 0.001 M, with other conditions being
standard, then Q = 0.1/0.001 = 100, and

Image ch2e59.jpg

Clearly, the reaction will now proceed in the direction of formation of G3P.

In a reaction A + B ⇌ C, in which two molecules combine to form a third, the equation for ΔG becomes

Image ch2e60.jpg

The direction of the reaction will shift more toward the right (toward formation of C) if either [A] or [B] is
increased.

Go to:

The ΔG°′ of a Reaction Can Be Calculated from Its Keq

A chemical mixture at equilibrium is already in a state of minimal free energy: no free energy is being
generated or released. Thus, for a system at equilibrium, we can write

Image ch2e61.jpg

At equilibrium the value of Q is the equilibrium constant Keq, so that

Image ch2e62.jpg

Expressed in terms of base 10 logarithms, this equation becomes

Image ch2e63.jpg

or
Image ch2e64.jpg

under standard conditions. Thus, if the concentrations of reactants and products at equilibrium (i.e., the
Keq) are determined, the value of ΔG°′ can be calculated. For example, we saw earlier that Keq equals
22.2 for the interconversion of glyceraldehyde 3-phosphate to dihydroxyacetone phosphate (G3P yz 
DHAP) under standard conditions. Substituting this value into Equation 2-9, we can easily calculate the
ΔG°′ for this reaction as −1840 cal/mol.

By rearranging Equation 2-9 and taking the antilogarithm, we obtain

Image ch2e65.jpg

From this expression, it is clear that if ΔG°′ is negative, then the exponent will be positive and hence Keq
will be greater than 1; that is, the formation of products from reactants is favored (Table 2-5).
Conversely, if ΔG°′ is positive, then the exponent will be negative and Keq will be less than 1.

Table 2-5. Values of ΔG°′ for Some Values of Keq.

Table 2-5

Values of ΔG°′ for Some Values of Keq.

Although a chemical equilibrium appears to be unchanging and static, it is actually a dynamic state. The
forward and the reverse reactions proceed at exactly the same rate, thereby canceling each other out.
As noted earlier, when an enzyme or some other catalyst speeds up a reaction, it also speeds up the
reverse reaction; thus equilibrium is reached sooner than it is when the reaction is not catalyzed.
However, the equilibrium constant and ΔG°′ of a reaction are the same in the presence and absence of a
catalyst.

Go to:

Cells Must Expend Energy to Generate Concentration Gradients

A cell must often accumulate chemicals, such as glucose and K+ ions, in greater concentrations than
exist in its environment. Consequently, the cell must transport these chemicals against a concentration
gradient. To find the amount of energy required to transfer 1 mole of a substance from outside the cell
to inside the cell, we use Equation 2-8 relating ΔG to the concentration of reactants and products.
Because this simple transport reaction does not involve making or breaking covalent bonds and no heat
is taken up or released, the ΔG°′ is 0. Thus Equation 2-8 becomes

Image ch2e66.jpg
where C2 is the initial concentration of a substance inside the cell and C1 is its concentration outside the
cell. If the ratio of C2 to C1 is 10, then at 25 °C, ΔG = RT ln 10 =  +1.36 kcal per mole of substance
transported. Such calculations assume that a molecule of a given substance inside a cell is identical with
a molecule of that substance outside and that the substance is not sequestered, bound, or chemically
changed by the transport.

Since the “uphill” transport of molecules against a concentration gradient (C2 > C1) has a positive ΔG, it
clearly cannot take place spontaneously. To occur, such transport requires the input of cellular chemical
energy, which often is supplied by the hydrolysis of ATP (Chapter 15). Conversely, when a substance
moves down its concentration gradient (C1 > C2) in crossing a membrane, ΔG has a negative value and
the transport can be coupled to a reaction that has a positive ΔG, say, the movement of another
substance uphill across a membrane.

Go to:

Many Cellular Processes Involve Oxidation-Reduction Reactions

Many chemical reactions result in the transfer of electrons from one atom or molecule to another; this
transfer may or may not accompany the formation of new chemical bonds. The loss of electrons from an
atom or a molecule is called oxidation, and the gain of electrons by an atom or a molecule is called
reduction. Because electrons are neither created nor destroyed in a chemical reaction, if one atom or
molecule is oxidized, another must be reduced. For example, oxygen draws electrons from Fe2+
(ferrous) ions to form Fe3+ (ferric) ions, a reaction that occurs as part of the process by which
carbohydrates are degraded in mitochondria. Each oxygen atom receives two electrons, one from each
of two Fe2+ ions:

Image ch2e67.jpg

Thus Fe2+ is oxidized, and O2 is reduced. Oxygen similarly accepts electrons in many oxidation reactions
in aerobic cells.

The transformation of succinate into fumarate is another oxidation reaction that takes place during
carbohydrate breakdown in mitochondria. In this reaction, succinate loses two hydrogen atoms, which is
equivalent to a loss of two protons and two electrons (Figure 2-23). Protons are soluble in aqueous
solutions (as H3O+), but electrons are not and must be transferred directly from one atom or molecule
to another. The electrons lost from succinate in its conversion to fumarate are transferred to flavin
adenine dinucleotide (FAD), which is reduced to FADH2. Many biologically important oxidation and
reduction reactions involve the removal or the addition of hydrogen atoms (protons plus electrons)
rather than the transfer of isolated electrons.

Figure 2-23. Succinate is converted to fumarate by the loss of two electrons and two protons.

Figure 2-23
Succinate is converted to fumarate by the loss of two electrons and two protons. This oxidation reaction,
which occurs in mitochondria as part of the citric acid cycle, is coupled to reduction of FAD to FADH2.

Standard Reduction Potentials

To describe oxidation-reduction reactions, such as the reaction of ferrous ion (Fe2+) and oxygen (O2), it
is easiest to divide them into two half-reactions:

Image ch2e68.jpg

In this case, the reduced oxygen (O2−) readily reacts with two protons to form one water molecule:

Image ch2e69.jpg

Thus if we add two protons to each side of the equation for the half-reaction for reduction of O2, the
half-reaction can be rewritten as

Image ch2e70.jpg

The readiness with which an atom or a molecule gains an electron is its reduction potential E. Reduction
potentials are measured in volts (V) from an arbitrary zero point set at the reduction potential of the
following half-reaction under standard conditions (25 °C, 1 atm, and reactants at 1 M):

Image ch2e71.jpg

The value of E for a molecule or an atom under standard conditions is its standard reduction potential, E
′0(Table 2-6). Standard reduction potentials may differ somewhat from those found under the
conditions in a cell, because the concentrations of reactants in a cell are not 1 M. A positive reduction
potential means that a molecule or ion (say, Fe3+) has a higher affinity for electrons than the H+ ion
does in the standard reaction. A negative reduction potential means that a substance — for example,
acetate (CH3COO−) in its reduction to acetaldehyde (CH3CHO) — has a lower affinity for electrons. In an
oxidation-reduction reaction, electrons move spontaneously toward atoms or molecules having more
positive reduction potentials. In other words, a compound having a more negative reduction potential
(or more positive oxidation potential) can reduce — or transfer electrons to — one having a more
positive reduction potential.

Table 2-6. Values of the Standard Reduction Potential E ′0 and Standard Free Energy ΔG°′ for Selected
Oxidation-Reduction Reactions (pH 7.0, 25 °C).

Table 2-6

Values of the Standard Reduction Potential E ′0 and Standard Free Energy ΔG°′ for Selected Oxidation-
Reduction Reactions (pH 7.0, 25 °C).
The Relationship between Changes in Free Energy and Reduction Potentials

In an oxidation-reduction reaction, the total voltage change (change in electric potential) ΔE is the sum
of the voltage changes (reduction potentials) of the individual oxidation or reduction steps. Because all
forms of energy are interconvertible, we can express ΔE as a change in chemical free energy (ΔG). The
charge in 1 mole (6.02 × 1023) of electrons is 96,500 coulombs (96,500 joules per volt), a quantity
known as the Faraday constant (ℱ) after British physicist Michael Faraday (1791 – 1867). The following
formula shows the relationship between free energy and reduction potential:

Image ch2e72.jpg

or

Image ch2e73.jpg

where n is the number of electrons transferred and 4.184 is the factor used to convert joules into
calories. Note that an oxidation-reduction reaction with a positive ΔE value will have a negative ΔG and
thus will tend to proceed from left to right.

The reduction potential is customarily used to describe the electric energy change that occurs when an
atom or a molecule gains an electron. In an oxidation-reduction reaction, we also use the oxidation
potential — the voltage change that takes place when an atom or molecule loses an electron — which is
simply the negative of the reduction potential:

Image ch2e74.jpg

The voltage change in a complete oxidation-reduction reaction, in which one molecule is reduced and
another is oxidized, is simply the sum of the oxidation potential and the reduction potential of the two
partial oxidation and reduction reactions, respectively. Consider, for example, the change in electric
potential (and, correspondingly, in standard free energy) when succinate is oxidized by oxygen:

Image ch2e75.jpg

In this case, the partial reactions are

Image ch2e76.jpg

The overall reaction has a positive ΔE′0 or, equivalently, a negative ΔG°′ and thus, under standard
conditions, will tend to occur from left to right.

Go to:

An Unfavorable Chemical Reaction Can Proceed If It Is Coupled with an Energetically Favorable Reaction
Many chemical reactions in cells are energetically unfavorable (ΔG > 0) and will not proceed
spontaneously. One example is the synthesis of small peptides (e.g., glycylalanine) or proteins from
amino acids. Cells are able to carry out a reaction that has a positive ΔG by coupling it to a reaction that
has a negative ΔG of larger magnitude, so that the sum of the two reactions has a negative ΔG. Suppose
that the reaction

Image ch2e77.jpg

has a ΔG°′ of +5 kcal/mol and that the reaction

Image ch2e78.jpg

has a ΔG°′ of −10 kcal/mol. In the absence of the second reaction, there would be much more A than B
at equilibrium. The occurrence of the second process, by which X becomes Y + Z, changes that outcome:
because it is such a favorable reaction, it will pull the first process toward the formation of B and the
consumption of A.

The ΔG°′ of the overall reaction will be the sum of the ΔG°′ values of each of the two partial reactions:

Image ch2e79.jpg

The overall reaction releases energy. In cells, energetically unfavorable reactions of the type A ⇌ B + X
are often coupled to the hydrolysis of the compound adenosine triphosphate (ATP), a reaction with a
negative change in free energy (ΔG°′ = −7.3 kcal/mol), so that the overall reaction has a negative ΔG°′.

Go to:

Hydrolysis of Phosphoanhydride Bonds in ATP Releases Substantial Free Energy

All cells extract energy from foods through a series of reactions that exhibit negative free-energy
changes; plant cells also can extract energy from absorbed light. In both cases, much of the free energy
is not allowed to dissipate as heat but is captured in chemical bonds formed by other molecules for use
throughout the cell. In almost all organisms, the most important molecule for capturing and transferring
free energy is adenosine triphosphate, or ATP (Figure 2-24).

Figure 2-24. In adenosine triphosphate (ATP), two high-energy phosphoanhydride bonds (red) link the
three phosphate groups.

Figure 2-24

In adenosine triphosphate (ATP), two high-energy phosphoanhydride bonds (red) link the three
phosphate groups.
The useful free energy in an ATP molecule is contained in phosphoanhydride bonds, which are formed
from the condensation of two molecules of phosphate by the loss of water:

Image ch2fu3.jpg

An ATP molecule has two phosphoanhydride bonds and is often written as adenosine – p~p~p, or simply
Ap~p~p, where p stands for a phosphate group and ~ denotes a high-energy bond.

Hydrolysis of a phosphoanhydride bond in each of the following reactions has a highly negative ΔG°′ of
about −7.3 kcal/mol:

Image ch2e80.jpg

In these reactions, Pi stands for inorganic phosphate and PPi for inorganic pyrophosphate, two
phosphate groups linked by a phosphoanhydride bond. As the top two reactions show, the removal of a
phosphate or a pyrophosphate group from ATP leaves adenosine diphosphate (ADP) or adenosine
monophosphate (AMP), respectively.

The phosphoanhydride bond is an ordinary covalent bond, but it releases about 7.3 kcal/mol of free
energy (under standard biochemical conditions) when it is broken. In contrast, hydrolysis of the
phosphoester bond in AMP, forming inorganic phosphate and adenosine, releases only about 2 kcal/mol
of free energy. Phosphoanhydride bonds commonly are termed “high-energy” bonds, even though the
ΔG°′ for the reaction of succinate with oxygen is much higher (−37 kcal/mol).

Cells can transfer the free energy released by the hydrolysis of phosphoanhydride bonds to other
molecules. This transfer supplies cells with enough free energy to carry out reactions that would
otherwise be unfavorable. For example, if the reaction

Image ch2e81.jpg

is energetically unfavorable (ΔG > 0), it can be made favorable by linking it to the hydrolysis of the
terminal phosphoanhydride bond in ATP. Some of the energy in this phosphoanhydride bond is used to
transfer a phosphate group to one of the reactants, forming a phosphorylated intermediate, B~p. The
intermediate thus has enough free energy to react with C, forming D and free phosphate:

Image ch2e82.jpg

Thus, the overall reaction is


Image ch2e83.jpg

which is energetically favorable. Chapter 4 illustrates in detail how the hydrolysis of ATP is coupled to
protein formation from amino acids; in the above example B and C would represent amino acids and D a
dipeptide. Cells keep the ratio of ATP to ADP and AMP high, often as high as 10:1. Thus reactions in
which the terminal phosphate group of ATP is transferred to another molecule will be driven even
further along.

As shown in Table 2-7, the ΔG°′ for hydrolysis of a phosphoanhydride bond in ATP (−7.3 kcal/mol) is
about twice the ΔG°′ for hydrolysis of a phosphoester bond, such as that in glucose 6-phosphate (−3.3
kcal/mol). A principal reason for this difference is that ATP and its hydrolysis products ADP and Pi are
highly charged at neutral pH. Three of the four ionizable protons in ATP are fully dissociated at pH 7.0,
and the fourth, with a pKa of 6.95, is about 50 percent dissociated. The closely spaced negative charges
in ATP repel each other strongly. When the terminal phosphoanhydride bond is hydrolyzed, some of this
stress is removed by the separation of the hydrolysis products ADP3− and HPO42−; that is, the
separated negatively charged ADP3− and HPO42− will tend not to recombine to form ATP. In glucose 6-
phosphate, by contrast, there is no charge repulsion between the phosphate group and the carbon atom
to which it is attached. One of the hydrolysis products, glucose, is uncharged and will not repel the
negatively charged HPO42− ion; thus there is less resistance to the recombination of glucose and
HPO42− to form glucose 6-phosphate.

Table 2-7. Values of ΔG°′ for the Hydrolysis of Various Biologically Important Phosphate Compounds*.

Table 2-7

Values of ΔG°′ for the Hydrolysis of Various Biologically Important Phosphate Compounds*.

Many other bonds — particularly those between a phosphate group and some other substance — have
the same high-energy character as phosphoanhydride bonds. The phosphoanhydride bond of ATP is not
the most or the least energetic of these bonds (see Table 2-7). The preeminent role of ATP in capturing
and transferring free energy within cells represents a compromise. The free energy of hydrolysis of ATP
is sufficiently great that reactions in which the terminal phosphate group is transferred to another
molecule have a substantially negative ΔG°′. However, if hydrolysis of this phosphoanhydride bond
liberated considerably more free energy than it does, cells might require too much energy to form this
bond in the first place. In other words, many reactions in cells release enough energy to form ATP, and
hydrolysis of ATP releases enough energy to drive many of the cell’s energy-requiring reactions and
processes.

Go to:

ATP Is Used to Fuel Many Cellular Processes

If the terminal phosphoanhydride bond of ATP were to rupture by hydrolysis to produce ADP and Pi,
energy would be released in the form of heat. However, cells contain various enzymes that can couple
ATP hydrolysis to other reactions, so that much of the released energy is converted to more useful forms
(Figure 2-25). For instance, cells use energy from ATP to synthesize macromolecules (proteins, nucleic
acids, and polysaccharides) and many types of small molecules. The hydrolysis of ATP also supplies the
energy needed to move individual cells from one location to another, to contract muscle cells, and to
transport molecules into or out of the cell, usually against a concentration gradient. Gradients of ions,
such as Na+ and K+, across a cellular membrane are produced by the action of membrane-embedded
enzymes, called ion pumps, that couple the hydrolysis of ATP to the “uphill” movement of ions. The
resulting ion concentration gradients are responsible for the generation of an electric potential across
the membrane. This potential is the basis for the electric activity of cells and, in particular, for the
conduction of impulses by nerves.

Figure 2-25. The ATP cycle.

Figure 2-25

The ATP cycle. ATP is formed from ADP and Pi by photosynthesis in plants and by the metabolism of
energy-rich compounds in most cells. The hydrolysis of ATP to ADP and Pi is linked to many key cellular
functions; the free energy released by the breaking (more...)

Clearly, to continue, functioning cells must constantly replenish their ATP supply. The ultimate energy
source for formation of high-energy bonds in ATP and other compounds in nearly all cells is sunlight.
Plants and microorganisms trap the energy in light through photosynthesis. In this process, chlorophyll
pigments absorb the energy of light, which is then used to synthesize ATP from ADP and Pi. Much of the
ATP produced in photosynthesis is used to help convert carbon dioxide to six-carbon sugars such as
fructose and glucose:

Image ch2e84.jpg

Additional energy is used to convert hexoses into the disaccharide sucrose and polysaccharides. In
animals, the free energy in sugars and other molecules derived from food is released in the process of
respiration. All synthesis of ATP in animal cells and in nonphotosynthetic microorganisms results from
the chemical transformation of energy-rich dietary or storage molecules. We discuss the mechanisms of
photosynthesis and cellular respiration in Chapter 16.

As noted earlier, glucose is a major source of energy in most cells. When 1 mole (180 g) of glucose reacts
with oxygen under standard conditions according to the following reaction, 686 kcal of energy is
released:

Image ch2e85.jpg
If glucose is simply burned in air, all this energy is released as heat. By an elaborate set of enzyme-
catalyzed reactions, cells couple the metabolism of 1 molecule of glucose to the synthesis of as many as
36 molecules of ATP from 36 molecules of ADP:

Image ch2e86.jpg

Because formation of one high-energy phosphoanhydride bond in ATP, from Pi and ADP, requires an
input of 7.3 kcal/mol, about 263 kcal of energy (36 × 7.3) is conserved in ATP per mole of glucose
metabolized (an efficiency of 263/686, or about 38 percent). This type of cellular metabolism is termed
aerobic because it is dependent on the oxygen in the air. Aerobic catabolism (degradation) of glucose is
found in all higher plant and animal cells and in many bacterial cells.

The overall reaction of glucose respiration

Image ch2e87.jpg

is the reverse of the photosynthetic reaction in which six-carbon sugars are formed

Image ch2e88.jpg

The latter reaction requires energy from light, whereas the former releases energy. Respiration and
photosynthesis are the two major processes constituting the carbon cycle in nature: sugars and oxygen
produced by plants are the raw materials for respiration and the generation of ATP by plant and animal
cells alike; the end products of respiration, CO2 and H2O, are the raw materials for the photosynthetic
production of sugars and oxygen. The only net source of energy in this cycle is sunlight. Thus, directly or
indirectly, light energy captured in photosynthesis is the source of chemical energy for almost all cells.

The exceptions to this are certain microorganisms that exist in deep ocean vents where sunlight is
completely absent. These unusual bacteria derive the energy for converting ADP and Pi into ATP from
the oxidation of reduced inorganic compounds present in the dissolved vent gas that originates in the
center of the earth. Unfortunately, little is yet known about the biology of these organisms.

Go to:

SUMMARY

 The change in free energy ΔG is the most useful measure for predicting the direction of chemical
reactions in biological systems. Chemical reactions tend to proceed in the direction for which ΔG is
negative.

 The ΔG of a reaction depends on the change in enthalpy ΔH (sum of bond energies), the change in
entropy ΔS (the randomness of molecular motion), and the temperature T: ΔG = ΔH − T ΔS.
 The standard free-energy change ΔG°′ equals −2.3 RT log Keq. Thus the value of ΔG°′ can be calculated
from the experimentally determined concentrations of reactants and products at equilibrium.

 The tendency of an atom or molecule to gain electrons is its reduction potential E, which is measured in
volts. The tendency to lose electrons is the oxidation potential, which has the same magnitude but
opposite sign as the reduction potential for the reverse reaction.

 Oxidation and reduction reactions always occur in pairs. The ΔE for an oxidation-reduction reaction is
the sum of the oxidation potential and the reduction potential of the two partial reactions. Oxidation-
reduction reactions with a positive ΔE have a negative ΔG and thus tend to proceed spontaneously.

 A chemical reaction having a positive ΔG can proceed if it is coupled with a reaction having a negative
ΔG of larger magnitude.

 Many energetically unfavorable cellular reactions are fueled by hydrolysis of one or both of the two
phosphoanhydride bonds in ATP.

 Directly or indirectly, light energy captured by photosynthesis in plants and photosynthetic bacteria is
the ultimate source of chemical energy for almost all cells.

You might also like