29.03.2013 Views

Lecture-Notes (Thermodynamics) - niser

Lecture-Notes (Thermodynamics) - niser

Lecture-Notes (Thermodynamics) - niser

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Chapter 4<br />

Entropy and second law of<br />

thermodynamics<br />

4.1 Carnot cycle<br />

In a cyclic transformation, where the final state of a system is the same as the initial one,<br />

∆U = 0<br />

since the internal energy U is a state function. A reversible cyclic process can be represented<br />

by a closed loop in the P − V diagram (Fig. 4.1).<br />

Figure 4.1: Reversible cyclic process. The shaded area of the loop is the total work done<br />

by the system in one cycle.<br />

It follows from the first law of thermodynamics that, since ∆U = 0, the work done by the<br />

system during a cycle is equal to the heat absorbed, i. e.,<br />

<br />

−∆W = ∆Q = PdV = Area enclosed.<br />

⋆ In a cyclic process, work is converted into heat and in the end of each cycle the<br />

system returns to its original state. We will introduce the heat engine which is<br />

based on the properties of a cyclic process.<br />

25


26 CHAPTER 4. ENTROPY AND SECOND LAW OF THERMODYNAMICS<br />

Let us consider the operation of a heat engine in more detail (Fig. 4.2).<br />

Part of the heat that is transferred to the system<br />

from a heat bath with temperature T1, Q1, is converted<br />

into work, W, and the rest, Q2, is delivered<br />

to a second bath with T2 < T1 (condenser). Following<br />

the first law of thermodynamics,<br />

Q1 − |Q2| = |W |.<br />

Carnot process: This a reversible cycle process<br />

bounded by two isotherms and two adiabatic lines.<br />

The Carnot process can be realized with an arbitrary<br />

working substance, but we shall consider here<br />

an ideal gas (Fig. 4.3).<br />

Figure 4.2: Heat engine.<br />

Figure 4.3: Ideal gas as a working substance for the Carnot process.<br />

In the P − V diagram the Carnot process can be represented as shown in Fig. 4.4.<br />

Figure 4.4: The Carnot process in the P − V diagram.<br />

One Carnot cycle consists of four consecutive thermodynamic processes, which are<br />

1. A → B: isothermal expansion at T = T1; the volume changes as<br />

VA → VB,<br />

while heat Q1 is absorbed from a bath and the system performs work.


4.2. SECOND LAW OF THERMODYNAMICS 27<br />

2. B → C: adiabatic expansion, during which<br />

and<br />

while δQ = 0.<br />

T1 → T2<br />

VB → VC,<br />

3. C → D: isothermal compression at T = T2; the system ejects heat Q2 to the bath<br />

(Q2 < 0 is a convention).<br />

4. D → A: adiabatic compression, during which<br />

T2 → T1,<br />

VD → VA,<br />

δQ = 0.<br />

In one cycle operation, the system receives an amount of heat Q1 from a hot reservoir,<br />

performs work, and rejects ”waste heat” Q2 to a cold reservoir.<br />

From the first law of thermodynamics we have:<br />

<br />

0 = dU = (δQ + δW) = Q + W = Q1 + Q2 + W,<br />

where −W is the work performed by the system, equal to the area enclosed in the loop<br />

(shaded area in Fig. 4.4).<br />

The efficiency of the Carnot engine is defined as<br />

η ≡<br />

performed work<br />

absorbed heat<br />

= −W<br />

Q1<br />

= Q1 + Q2<br />

Q1<br />

= Q1 − |Q2|<br />

Q1<br />

η is 100% if there is no waste heat (Q2 = 0). However, we will see that this is impossible<br />

due to the second law of thermodynamics.<br />

4.2 Second law of thermodynamics<br />

Definition by Clausius:<br />

” There is no thermodynamic transformation whose sole effect is to deliver heat<br />

from a reservoir of lower temperature to a reservoir of higher temperature.”<br />

Summary: heat does not flow upwards.<br />

Definition by Kelvin:<br />

”There is no thermodynamic transformation whose sole effect is to extract heat<br />

from a reservoir and convert it entirely to work”.<br />

Summary: a perpetuum mobile of second type does not exist.<br />

.


28 CHAPTER 4. ENTROPY AND SECOND LAW OF THERMODYNAMICS<br />

In order to prove that both definition are equivalent, we will<br />

show that the falsehood of one implies the falsehood of the<br />

other. For that purpose, we consider two heat reservoirs with<br />

temperatures T1 and T2 with T1 > T2.<br />

(1) If Kelvin’s statement were false, we could extract heat<br />

from T2 and convert it entirely to work. We could then<br />

convert the work back to heat entirely and deliver it to T1<br />

(there is no law against this) (Fig. 4.5). Thus, Clausius’<br />

statement would be negated.<br />

Figure 4.5: Q2 = W =<br />

Q1 process.<br />

(2) If Clausius’ statement were false, we could let an amount<br />

of heat Q1 flow from T2 to T1 (T2 < T1). Then, we could<br />

connect a Carnot engine between T1 and T2 such as to<br />

extract Q1 from T1 and return an amount |Q2| < Q1 back to T2. The net work<br />

output of such an engine would be |Q1| − |Q2| > 0, which would mean that an<br />

amount of heat |Q1| − |Q2| is converted into work, without any other effect. This<br />

would contradict Kelvin’s statement.<br />

From the microscopic point of view<br />

- heat transfer is an exchange of energy due to the random motion of atoms;<br />

- work’s performance requires an organized action of atoms.<br />

In these terms, heat being converted entirely into work means chaos changing spontaneously<br />

to order, which is a very improbable process.<br />

→ Usually,<br />

- one configuration corresponds to order;<br />

- many configurations correspond to chaos.


4.3. ABSOLUTE TEMPERATURE 29<br />

4.3 Absolute temperature<br />

In order to introduce the concept of absolute temperature, let us discuss the consequences<br />

of the second law of thermodynamics.<br />

i) The second law of thermodynamics implies that a Carnot engine cannot be 100%<br />

efficient since otherwise all heat absorbed from a warm reservoir would be converted<br />

into work in one cycle.<br />

ii) No engine working between two given temperatures can be more efficient than a<br />

Carnot engine.<br />

A Carnot engine is a reversible engine between two baths. The previous statement means<br />

as well that an irreversible engine cannot be more efficient than a reversible one. In order<br />

to show this, we consider two engines C and X (with X not necessarily a Carnot one)<br />

working between baths at T1 (warm) and T2 (cold). We run the Carnot engine C in<br />

reverse, as a refrigerator ¯ C, and feed the work output of X to ¯ C (Fig. 4.6).<br />

Figure 4.6: Linked engine X and Carnot refrigerator ¯ C.<br />

The total work output of such a system is<br />

Wtot = (|Q ′ 1| − |Q ′ 2|) − (|Q1| − |Q2|) .<br />

If we adjust the two engines so that |Q ′ 1 | = |Q1|, no net heat is extracted from the heat<br />

bath at T1. In this case, an amount of heat |Q2| − |Q ′ 2 | is extracted from the heat bath<br />

at T2 and converted entirely to work, with no other effect. This would violate the second<br />

law of thermodynamics unless<br />

|Q2| ≤ |Q ′ 2 |.<br />

We divide this inequality by |Q1| and, using the fact that |Q1| = |Q ′ 1 |, get<br />

and<br />

|Q2|<br />

|Q1| ≤ |Q′ 2 |<br />

|Q ′ 1|<br />

1 − |Q2|<br />

|Q1| ≥ 1 − |Q′ 2|<br />

.<br />

|<br />

|Q ′ 1


30 CHAPTER 4. ENTROPY AND SECOND LAW OF THERMODYNAMICS<br />

From this follows that the efficiencies of the engines compare as<br />

ηC ≥ ηX .<br />

Consequence: all Carnot engines have the same efficiency since X can be, as a special<br />

case, a Carnot engine.<br />

⇒ The Carnot engine is universal, i.e., it depends only on the temperatures involved<br />

and not on the working substance.<br />

Definition: absolute temperature θ of a heat reservoir (bath) is defined such that the ratio<br />

of the absolute temperatures of two reservoirs is given by<br />

θ2<br />

θ1<br />

≡ |Q2|<br />

|Q1|<br />

−Q2<br />

= 1 − η = ,<br />

where η is the efficiency of a Carnot engine operating between the two reservoirs. Since<br />

|Q2| is strictly greater than zero, |Q2| > 0, according to the second law of thermodynamics,<br />

the absolute temperature is bounded from below:<br />

θ > 0 .<br />

⋆ This means that the absolute zero θ = 0 is a limiting value that can never be reached<br />

since this would violate the second law of thermodynamics.<br />

If an ideal gas is considered as a working substance in a Carnot engine, the absolute<br />

temperature coincides with the temperature of the ideal gas defined as<br />

T = PV<br />

NkB<br />

= θ.<br />

4.4 Temperature as integrating factor<br />

The absolute temperature T can be considered as an integrating factor that converts the<br />

inexact differential δQ into an exact differential δQ/T.<br />

Let us consider the following theorem stated by Clausius:<br />

⇒ ”In an arbitrary cyclic process P, the following inequality holds:<br />

<br />

P<br />

δQ<br />

T<br />

≤ 0;<br />

where the equality holds for P reversible.“<br />

Q1


4.4. TEMPERATURE AS INTEGRATING FACTOR 31<br />

Proof. Let us divide the cycle P into n segments so that on each segment<br />

its temperature Ti (i = 1, . . .,n) is constant. We consider now a reservoir<br />

at temperature T0 > Ti(∀i) (Fig. 4.7) and introduce Carnot engines<br />

between the reservoir at T0 and Ti.<br />

Figure 4.7: Cycle P connected to the reservoir at T0 via Carnot engines.<br />

For each Carnot engine,<br />

Wi + Q (0)<br />

i + QC i<br />

For cycle P,<br />

with<br />

Q (0)<br />

i<br />

T0<br />

+ QC i<br />

Ti<br />

= 0 (first law of thermodynamics) and<br />

= 0 (definition of absolute temperature).<br />

W +<br />

n<br />

Qi = 0 (first law of thermodynamics),<br />

i=1<br />

Qi = −Q C i .<br />

Then, the total heat absorbed from the reservoir at T0 is<br />

n<br />

n Q<br />

= −T0<br />

C i<br />

= T0<br />

Q (0)<br />

T =<br />

i=1<br />

Q (0)<br />

i<br />

and the work performed by the system is<br />

<br />

n<br />

WT = − W +<br />

=<br />

n<br />

Qi +<br />

i=1<br />

i=1<br />

n<br />

i=1<br />

Wi<br />

<br />

i=1<br />

Ti<br />

<br />

Q (0)<br />

<br />

i − Qi =<br />

n<br />

i=1<br />

n<br />

i=1<br />

Q (0)<br />

i<br />

Qi<br />

Ti<br />

= Q(0)<br />

T .


32 CHAPTER 4. ENTROPY AND SECOND LAW OF THERMODYNAMICS<br />

> 0, the combined machine would convert completely the heat from reservoir at<br />

T0 into mechanical work. This would violate Kelvin’s principle and therefore<br />

If Q (0)<br />

T<br />

i.e.,<br />

For n → ∞,<br />

The theorem of Clausius is proved.<br />

T0<br />

<br />

Q (0)<br />

T<br />

n<br />

i=1<br />

P<br />

δQ<br />

T<br />

≤ 0,<br />

Qi<br />

Ti<br />

≤ 0.<br />

≤ 0 .<br />

For a reversible process, we can revert the energy, which gives us<br />

<br />

δQ<br />

≥ 0.<br />

T<br />

This implies that<br />

for a reversible process.<br />

<br />

P<br />

P<br />

δQ<br />

T<br />

= 0<br />

Corollary: for a reversible open path P, the integral<br />

<br />

dQ<br />

T<br />

depends only on the end points and not on the particular path.<br />

4.5 Entropy<br />

Since<br />

dQ<br />

T<br />

= 0 for a reversible process, δQ<br />

T<br />

P<br />

dS = δQ<br />

T<br />

Eq. (4.1) implies that there exists a state function S.<br />

is an exact differential:<br />

. (4.1)<br />

Definition: The state function S (potential), whose differential is given as<br />

is called entropy.<br />

dS = δQ<br />

T


4.5. ENTROPY 33<br />

The entropy is defined up to an additive constant. The difference between entropies of<br />

any two states A and B is<br />

B<br />

δQ<br />

S(B) − S(A) ≡<br />

T ,<br />

where the integral extends along any reversible path connecting A and B, and the result<br />

of the integration is independent of the path.<br />

What happens when the integration is along an irreversible<br />

path? Since I − R is a cycle (see Fig. 4.8), it<br />

follows from Clausius’ theorem that<br />

⇒<br />

<br />

I<br />

<br />

δQ<br />

T ≤<br />

Therefore, in general<br />

B<br />

A<br />

I−R<br />

<br />

δQ<br />

T<br />

R<br />

δQ<br />

T<br />

δQ<br />

T<br />

≤ 0 ⇒<br />

= S(B) − S(A) .<br />

≤ S(B) − S(A) ,<br />

and the equality holds for a reversible process.<br />

A<br />

Figure 4.8: I − R (irreversiblereversible)<br />

cycle.<br />

In particular, for an isolated system, which does not exchange heat with a reservoir,<br />

δQ = 0 and therefore<br />

∆S ≥ 0 .<br />

This means that the entropy of an isolated system never decreases and remains constant<br />

during a reversible transformation.<br />

Note:<br />

i) The joint system of a system and its environment is called ”universe”. Defined in this<br />

way, the ”universe” is an isolated system and, therefore, its entropy never decreases.<br />

However, the entropy of a non-isolated system may decrease at the expense of the<br />

system’s environment.<br />

ii) Since the entropy is a state function, S(B) − S(A) is independent of the path, regardless<br />

whether it is reversible or irreversible. For an irreversible path, the entropy<br />

of the environment changes, whereas for a reversible one it does not.<br />

iii) Remember that the entropy difference<br />

B<br />

S(B) − S(A) =<br />

only when the path is reversible; otherwise the difference is larger that the integral.<br />

A<br />

δQ<br />

T


34 CHAPTER 4. ENTROPY AND SECOND LAW OF THERMODYNAMICS<br />

4.6 The energy equation<br />

We note that for a reversible process in a closed system the first law of thermodynamics<br />

can be written as<br />

TdS = dU − δW = dU + PdV.<br />

Let us consider T and V as independent variables (f.i., in a gas):<br />

S = S(T, V ),<br />

U = U(T, V ) → caloric equation of state,<br />

P = P(T, V ) → thermic equation of state.<br />

Since usually the internal energy U is not directly measurable, it is more practical to<br />

consider dS = δQ/T. We will make use of one of the heat equations derived in section<br />

3.5:<br />

Dividing both sides of eq. (4.2) by T we get<br />

On the other hand,<br />

<br />

∂U ∂U<br />

δQ = dT + + P dV<br />

∂T V ∂V T <br />

∂U<br />

= CV dT + + P dV (4.2)<br />

∂V<br />

δQ<br />

T<br />

= dS = CV<br />

T<br />

dS =<br />

dT + 1<br />

T<br />

T<br />

<br />

∂U<br />

+ P dV . (4.3)<br />

∂V T<br />

<br />

∂S ∂S<br />

dT + dV . (4.4)<br />

∂T V ∂V T<br />

Comparing eq. (4.3) and (4.4), we obtain the following expressions:<br />

<br />

∂S<br />

∂T V<br />

<br />

∂S<br />

∂V<br />

T<br />

= CV<br />

T ,<br />

= 1<br />

T<br />

<br />

∂U<br />

+ P .<br />

∂V T<br />

We use the commutativity of differentiation operations,<br />

to write<br />

1<br />

T<br />

∂<br />

∂V<br />

<br />

∂U<br />

∂T V T<br />

∂<br />

∂V<br />

∂S<br />

∂T<br />

T<br />

= ∂<br />

∂T<br />

∂S<br />

∂V ,<br />

<br />

∂ CV<br />

=<br />

∂V T<br />

∂<br />

<br />

1 ∂U<br />

∂T T ∂V T<br />

= − 1<br />

T 2<br />

<br />

∂U<br />

+ P +<br />

∂V<br />

1<br />

<br />

∂<br />

T ∂T<br />

<br />

+ P ,<br />

<br />

∂U<br />

∂V T V<br />

+<br />

<br />

∂P<br />

;<br />

∂T V


4.7. USEFUL COEFFICIENTS 35<br />

we cancel identical terms:<br />

and get<br />

<br />

∂U<br />

∂V T<br />

= T<br />

∂<br />

∂V<br />

<br />

∂P<br />

∂T V<br />

<br />

∂U<br />

=<br />

∂T<br />

∂U<br />

∂T∂V<br />

− P ⇒ energy equation. (4.5)<br />

In the energy equation, the derivative of the internal energy is written in terms of measurable<br />

quantities.<br />

Examples: for an ideal gas <br />

∂U<br />

∂V T<br />

= T nR<br />

V<br />

− P = 0,<br />

which means that the internal energy of the ideal gas is not dependent on its volume:<br />

Gay-Lussac experiment:<br />

4.7 Useful coefficients<br />

U = U(T) = CV T + const.<br />

U = 3<br />

2 NkBT → gas of monoatomic molecules,<br />

U = 5<br />

2 NkBT → gas of diatomic molecules.<br />

As<br />

<br />

we<br />

<br />

saw in the last section, the energy equation relates the experimentally inaccessible<br />

∂U<br />

∂V T to the experimentally accessible <br />

∂P<br />

∂T V . We will show now that <br />

∂P<br />

can be<br />

∂T V<br />

related to other thermodynamic coefficients as well. For that purpose, we will make use<br />

of the chain rule<br />

<br />

∂x ∂y ∂z<br />

= −1 ,<br />

∂y z ∂z x ∂x y<br />

(4.6)<br />

which can be derived by considering the differential of function f(x, y, z):<br />

df = ∂f ∂f ∂f<br />

dx + dy +<br />

∂x ∂y ∂z dz,<br />

<br />

∂x<br />

∂y<br />

<br />

∂y<br />

∂z<br />

z<br />

x<br />

= − ∂f/∂y<br />

∂f/∂x ,<br />

= − ∂f/∂z<br />

∂f/∂y ,


36 CHAPTER 4. ENTROPY AND SECOND LAW OF THERMODYNAMICS<br />

Using (4.6), we get<br />

<br />

∂P<br />

∂T V<br />

where<br />

= −<br />

α = 1<br />

<br />

∂V<br />

V ∂T P<br />

kT = − 1<br />

<br />

∂V<br />

V ∂P T<br />

kS = − 1<br />

<br />

∂V<br />

V ∂P<br />

<br />

∂z<br />

= −<br />

∂x y<br />

∂f/∂x<br />

∂f/∂z .<br />

1<br />

(∂T/∂V ) P (∂V/∂P) T<br />

S<br />

= − (∂V/∂T) P<br />

(∂V/∂P) T<br />

= α<br />

kT<br />

(coefficient of thermal expansion),<br />

(isothermal compressibility),<br />

(adiabatic compressibility).<br />

We substitute in<br />

<br />

∂U<br />

δQ = TdS = CV dT + + P dV<br />

∂V T<br />

expressions (4.7) and (4.5) and get the following useful relations.<br />

, (4.7)<br />

1) The absorbed heat is expressed in terms of directly measurable coefficients as<br />

where T and V are independent variables.<br />

2) If T and P are used as independent variables, then<br />

δQ = TdS = CV dT + α<br />

TdV , (4.8)<br />

kT<br />

TdS = CPdT − αTV dP (4.9)<br />

3) If V and P are used as independent variables, dT can be rewritten in terms of dV<br />

and dP as<br />

<br />

∂T ∂T<br />

dT = dV + dP =<br />

∂V P ∂P V<br />

1 kT<br />

+ dP .<br />

αV α<br />

The corresponding TdS equation is then<br />

TdS = CP<br />

dV +<br />

αV<br />

CPkT<br />

α<br />

<br />

− αTV dP . (4.10)<br />

There is also an important connection between CP and CV (which follows from eq.<br />

(4.8) and (4.9)):<br />

CP − CV = α2<br />

TV = −T<br />

kT<br />

<br />

∂V 2<br />

∂T P<br />

∂V<br />

∂P T<br />

> 0 . (4.11)


4.8. ENTROPY AND DISORDER 37<br />

By introducing the isochore pressure coefficient<br />

it is possible to rewrite CP − CV as<br />

Indeed,<br />

β = 1<br />

P<br />

α<br />

kT<br />

⇒ CP − CV = α2<br />

β = 1<br />

P<br />

4.8 Entropy and disorder<br />

kT<br />

<br />

∂P<br />

∂T V<br />

β 2 TV kTP 2 .<br />

⇒ α = βPkT ⇒<br />

TV = β2 P 2 k 2 T<br />

kT<br />

,<br />

TV = β 2 TV kTP 2 .<br />

We want now to establish a connection between thermodynamics and statistical mechanics,<br />

which we will be learning in the second half of the course.<br />

1. We give a definition of the multiplicity of a macrostate as<br />

⎛<br />

number of microstates<br />

multiplicity of<br />

= ⎝ that correspond to the<br />

a macrostate<br />

macrostate<br />

2. The entropy can be defined in terms of the multiplicity as<br />

S ≡ k ln (multiplicity) .<br />

We shall see in statistical mechanics that this definition is equivalent to the phenomenological<br />

concept we have learnt previously in this section.<br />

3. With the above definition of entropy at hand, we can formulate the second law of<br />

thermodynamics in this way:<br />

4.<br />

”If an isolated system of many particles is allowed to change, then, with large probability,<br />

it will evolve to the macrostate of largest entropy and will remain in that<br />

macrostate.”<br />

(energy input by heating)<br />

∆Sany system ≥ ,<br />

T<br />

where T is the temperature of the reservoir.<br />

5. If two macroscopic systems are in thermal equilibrium and in thermal contact, the<br />

entropy of the composite system equals the sum of the two individual entropies.<br />

⎞<br />

⎠.


38 CHAPTER 4. ENTROPY AND SECOND LAW OF THERMODYNAMICS<br />

6. Since entropy is proportional to the ”multiplicity of a macrostate”, it can be viewed<br />

as a ”measure of disorder”:<br />

Order Disorder<br />

⇓ ⇓<br />

strong correlation absence of correlation<br />

⇓ ⇓<br />

small multiplicity large multiplicity<br />

As an illustration, let us consider a class in the school, where all kids sit always in<br />

the same place. This situation can be defined as ”ordered” since there is a strong<br />

correlation between the kids’ positions. For example, if Peter sits in the first row,<br />

you know straight away who is sitting behind and besides him: there is only one<br />

possibility (multiplicity is very low).<br />

Looking now at a class in the kindergarten, we will find that the children there are<br />

all mixed, sitting and moving all the time. Now, if Peter sits in the first row, you<br />

have no idea where the other kids are sitting. There are no correlations between the<br />

kids’ positions and there are many options to have the children in the class. Such a<br />

”system” of kids is in a ”disordered state”, with very high multiplicity. The entropy<br />

of the kindergarten class is high!<br />

4.9 Third law of thermodynamics (Nernst law)<br />

While with the second law of thermodynamics we can only define the entropy (up to a<br />

constant), the third law of thermodynamics tells us how S(T) behaves at T → 0. Namely,<br />

it states that the entropy of a thermodynamic system at T = 0 is a universal constant,<br />

that we take to be 0:<br />

lim S(T) = 0.<br />

T →0<br />

S(T → 0) is independent from the values that the rest of variables take.<br />

Consequences:<br />

1) The heat capacities of all substances disappear at T = 0:<br />

lim<br />

T →0 CV<br />

<br />

∂S<br />

= lim T = 0,<br />

T →0 ∂T V<br />

lim<br />

T →0 CP<br />

<br />

∂S<br />

= lim T = 0.<br />

T →0 ∂T P<br />

Note that an ideal gas provides a contradiction to the previous statement since in<br />

that case CV = const and CP = const:<br />

CV = 3<br />

2 NkB,<br />

CP = CV + NkB = 5<br />

2 NkB,<br />

CP − CV = NkB = 0. (4.12)


4.9. THIRD LAW OF THERMODYNAMICS (NERNST LAW) 39<br />

2)<br />

But, for T → 0 the ideal gas is not anymore a realistic system because a real gas<br />

undergoes at low temperatures a phase transition - condensation.<br />

lim β = lim<br />

T →0 T →0<br />

3) The absolute T = 0 (zero point) is unattainable.<br />

1<br />

V<br />

<br />

∂V<br />

= 0<br />

∂T<br />

In order to show the validity of the last statement, let us analyze what happens when we<br />

are trying to reach low temperatures by subsequently performing adiabatic and isothermal<br />

transformations. If a gas is used as a working substance (Linde method), such a sequence<br />

of transformations looks in the P − V diagram as shown in Fig. 4.9.<br />

Figure 4.9: Linde method: sequence of adiabatic and isothermal transformations in a gas.<br />

Let us consider now the processes involved.<br />

A → B: isothermic compression.<br />

Work is performed on the gas and an amount of heat Q1 < 0 is given to the reservoir<br />

in a reversible process. As a result, ∆S = Q1<br />

T diminishes.<br />

B → C: adiabatic expansion.<br />

The gas performs work. Since δQ = 0, the entropy remains constant and the<br />

temperature diminishes.


40 CHAPTER 4. ENTROPY AND SECOND LAW OF THERMODYNAMICS<br />

In principle, by repeating this process many times, one could reach the T = 0 point.<br />

But, according to the third law of thermodynamics, all entropy curves must end at 0 for<br />

T → 0, which means that only through an infinite number of steps is it possible to reach<br />

T = 0.

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!