Lecture Notes Set 3b: Entropy and the 2 nd law

Similar documents
Lecture Notes Set 3a: Probabilities, Microstates and Entropy

Lecture Notes Set 4c: Heat engines and the Carnot cycle

Physics 172H Modern Mechanics

Chapter 20 The Second Law of Thermodynamics

Entropy, free energy and equilibrium. Spontaneity Entropy Free energy and equilibrium

Lecture 8. The Second Law of Thermodynamics; Energy Exchange

Lecture 8. The Second Law of Thermodynamics; Energy Exchange

Irreversible Processes

12 The Laws of Thermodynamics

PHYSICS 715 COURSE NOTES WEEK 1

Irreversible Processes

Irreversible Processes

Heat Machines (Chapters 18.6, 19)

Chapter 12. The Laws of Thermodynamics

Chapter 20. Heat Engines, Entropy and the Second Law of Thermodynamics. Dr. Armen Kocharian

Chapter 4 - Second Law of Thermodynamics

October 18, 2011 Carnot cycle - 1

Chapter 12. The Laws of Thermodynamics. First Law of Thermodynamics

Lecture 15. Available Work and Free Energy. Lecture 15, p 1

Thermodynamic entropy

Basic Thermodynamics. Prof. S. K. Som. Department of Mechanical Engineering. Indian Institute of Technology, Kharagpur.

Introduction to thermodynamics

Physics 207 Lecture 27. Lecture 26. Chapters 18, entropy and second law of thermodynamics Chapter 19, heat engines and refrigerators

Thermodynamics. 1.1 Introduction. Thermodynamics is a phenomenological description of properties of macroscopic systems in thermal equilibrium.

Lecture 2 Entropy and Second Law

Entropy in Macroscopic Systems

Spontaneity: Second law of thermodynamics CH102 General Chemistry, Spring 2012, Boston University

Adiabatic Expansion (DQ = 0)

Class 22 - Second Law of Thermodynamics and Entropy

Thermodynamics & Statistical Mechanics SCQF Level 9, U03272, PHY-3-ThermStat. Thursday 24th April, a.m p.m.

CARNOT CYCLE = T = S ( U,V )

Lecture Notes 2014March 13 on Thermodynamics A. First Law: based upon conservation of energy

Stuff 1st Law of Thermodynamics First Law Differential Form Total Differential Total Differential

Chapter 16 Thermodynamics

Chapter 12 Thermodynamics

Lecture 9 Examples and Problems

Random arrang ement (disorder) Ordered arrangement (order)

(prev) (top) (next) (Throughout, we will assume the processes involve an ideal gas with constant n.)

X α = E x α = E. Ω Y (E,x)

Heat What is heat? Work = 2. PdV 1

1. Second Law of Thermodynamics

213 Midterm coming up

Physics 121, April 29, The Second Law of Thermodynamics.

Lecture 10: Heat Engines and Reversible Processes

THE SECOND LAW OF THERMODYNAMICS. Professor Benjamin G. Levine CEM 182H Lecture 5

Section 3 Entropy and Classical Thermodynamics

Imperial College London BSc/MSci EXAMINATION May 2008 THERMODYNAMICS & STATISTICAL PHYSICS

S = S(f) S(i) dq rev /T. ds = dq rev /T

The Second Law of Thermodynamics

Let s start by reviewing what we learned last time. Here is the basic line of reasoning for Einstein Solids

Chapter 12- The Law of Increasing Disorder

Physics 1501 Lecture 37

Thermodynamic Systems, States, and Processes

DAY 28. Summary of Primary Topics Covered. The 2 nd Law of Thermodynamics

11/29/2017 IRREVERSIBLE PROCESSES. UNIT 2 Thermodynamics: Laws of thermodynamics, ideal gases, and kinetic theory

Chapter 11 Heat Engines and The Second Law of Thermodynamics

Basic Thermodynamics Prof. S. K. Som Department of Mechanical Engineering Indian Institute of Technology, Kharagpur

PY2005: Thermodynamics

What is thermodynamics? and what can it do for us?

Chemistry 163B Heuristic Tutorial Second Law, Statistics and Entropy

THERMODYNAMICS. Zeroth law of thermodynamics. Isotherm

Thermodynamics General

The First Law of Thermodynamics

How to please the rulers of NPL-213 the geese

ChE 503 A. Z. Panagiotopoulos 1

(Refer Slide Time: 00:00:43 min) Welcome back in the last few lectures we discussed compression refrigeration systems.

Thermodynamics: Entropy

Handout 12: Thermodynamics. Zeroth law of thermodynamics

Second Law of Thermodynamics: Concept of Entropy. While the first law of thermodynamics defines a relation between work and heat, in terms of

THERMODYNAMICS b) If the temperatures of two bodies are equal then they are said to be in thermal equilibrium.

a. 4.2x10-4 m 3 b. 5.5x10-4 m 3 c. 1.2x10-4 m 3 d. 1.4x10-5 m 3 e. 8.8x10-5 m 3

Physics 9 Friday, November 2, 2018

8 Lecture 8: Thermodynamics: Principles

The need for something else: Entropy

Handout 12: Thermodynamics. Zeroth law of thermodynamics

Entropy and the Second and Third Laws of Thermodynamics

The Second Law of Thermodynamics

More Thermodynamics. Specific Specific Heats of a Gas Equipartition of Energy Reversible and Irreversible Processes

Chapter 20 Entropy and the 2nd Law of Thermodynamics

THERMODYNAMICS SSC-JE STAFF SELECTION COMMISSION MECHANICAL ENGINEERING STUDY MATERIAL THERMODYNAMICS THERMODYNAMICS THERMODYNAMICS

THERMODYNAMICS Lecture 5: Second Law of Thermodynamics

Thermodynamics: More Entropy

Engineering Thermodynamics. Chapter 5. The Second Law of Thermodynamics

1985B4. A kilogram sample of a material is initially a solid at a temperature of 20 C. Heat is added to the sample at a constant rate of 100

CHAPTER - 12 THERMODYNAMICS

Lecture Outline Chapter 18. Physics, 4 th Edition James S. Walker. Copyright 2010 Pearson Education, Inc.

Classical thermodynamics

Entropy. Physics 1425 Lecture 36. Michael Fowler, UVa

8.21 The Physics of Energy Fall 2009

AME 436. Energy and Propulsion. Lecture 7 Unsteady-flow (reciprocating) engines 2: Using P-V and T-s diagrams

Distinguish between an isothermal process and an adiabatic process as applied to an ideal gas (2)

Chapter 19. Heat Engines

Chemistry 163B Heuristic Tutorial Second Law, Statistics and Entropy falling apples excited state ground state chemistry 163B students

Last Name: First Name ID

Lecture 25 Goals: Chapter 18 Understand the molecular basis for pressure and the idealgas

dt T du T = C V = Nk ln = Nk ln 1 + V ]

Quiz 3 for Physics 176: Answers. Professor Greenside

CHEM Introduction to Thermodynamics Fall Entropy and the Second Law of Thermodynamics

The goal of thermodynamics is to understand how heat can be converted to work. Not all the heat energy can be converted to mechanical energy

Statistical Physics. The Second Law. Most macroscopic processes are irreversible in everyday life.

Transcription:

Lecture Notes Set 3b: Entropy and the 2 nd law

3.5 Entropy and the 2 nd law of thermodynamics The st law of thermodynamics is simply a restatement of the conservation of energy principle and may be concisely written as: U = Q + W where U, Q, and W are the change in internal energy of the system, the heat output or input, and the work done (or on) the system respectively. Remarkably, Carnot (who is pictured to the right) developed the 2 nd law before the conservation of energy principle had been put forward. We ll encounter very many equivalent descriptions of the 2 nd law in the coming sections (and we ll also not infrequently encounter the work and genius of Carnot). For now, we ll start with the following statements: The entropy of a thermally isolated system increases in any reversible process and is unaltered in a reversible process (p.79, Thermal Physics, CB Finn) If a closed system is not in equilibrium, the most probable consequence is that the entropy will increase (p.354, Matter and Interactions Vol. I, Chabay and Sherwood). A more succinct statement of the second law is as follows: a closed system will tend towards maximum entropy. There are very many pitfalls associated with unclear language when considering entropy and the 2 nd law. Sloppy thinking and poorly worded statements of the laws of thermodynamics (which I ll do my very best to avoid) have led to many abuses of the concept of entropy some of these are described briefly in the final section of this set of the lecture notes. Before proceeding, we ll define some of the terms in the statements above: thermally isolated system : by this we mean a system surrounded by adiabatic walls there is no heat flow into or out of the system; closed system : this means that there is no energy (or matter) flow into or out of the system the Universe as a whole is the archetypal example of a closed system but small portions of the Universe can also be considered closed systems under certain circumstances (e.g. when isolated using adiabatic barriers). will tend towards maximum entropy : this is a very important phrase - the tendency is for entropy to increase, but echoing our discussion of Boltzmann factors in Set 2 of the notes, the system can be constrained from reaching the maximum entropy state the microstates may be inaccessible. (Indeed, we ll see later on in this section that S = k ln W and the Boltzmann factor expression given in Section 2 are just different statements of the same law).

The concepts of reversibility and irreversibility will be encountered a considerable number of times in the remainder of the module. We return to an introductory discussion of reversibility in Section 3.6. A Q Adiabatic wall B Fig. 3.2 A small amount of heat is transferred from Block A to Block B which are both contained within an adiabatic enclosure. The temperature of block A is lowered whereas that of Block B is increased. The number of accessible microstates in system B is increased whereas there are fewer accessible microstates in Block A following the energy transfer. Hence, although the entropy of Block B increases that of Block A is reduced. One common misconception with regard to the 2 nd law is that the entropy of an object always increases. That this is incorrect can be shown by considering the system shown in Fig. 3.2. As heat is transferred from Block A to Block B, the temperature of Block A is reduced and that of Block B is increased. Hence, the entropy of Block A is reduced in this case. However, the net entropy of the closed system is increased. (You ll quantitatively estimate entropy changes for a similar system in Coursework Set 6). The net increase in entropy arises because the most probable energy distribution is that which maximizes the total number of microstates (sections 3.3 and 3.4). It is worth noting at this point that although we re concerned with energy distributions, the concept of microstates is equally applicable to spatial distributions of particles. The analysis of diffusion processes in Set 2c of the notes illustrated that particles diffuse from regions of high to low concentration, so that at equilibrium no concentration gradients exist and the particles have spread out to fill the available space. The equilibrium spatial distribution is associated with a uniform, homogeneous spread of particles. This homogeneous spread is analogous to the uniform distribution of energy quanta observed for a system at thermal equilibrium. In both cases the uniform distribution of particles arises because the entropy for that macrostate far outweighs that of other less homogeneous arrangements. For the same reason that we don t observe cold objects becoming colder when placed in contact with a hot object, the odds against particles moving from a homogeneous distribution to an inhomogeneous, ordered arrangement are such that on average we d have to wait for a period very many times the age of our Universe to see such an arrangement appear for a mole of atoms. Particles spread out to fill the available space uniformly because the largest number of microstates (by far) is associated with this disordered macrostate.

NOTES

3.6 Reversibility One of the statements of the 2 nd law given above notes that the entropy of a thermally isolated system is unaltered in a reversible process. What precisely is meant by reversible in the context of thermodynamics? In a reversible process the system must be capable of being returned to its original state with no other change in the surroundings. The classic example of a reversible process involves the pistongas system shown in Fig. 3.3. Here a gas undergoes an infinitesimally small compression due to an infinitesimally small weight, dm, which is placed on the frictionless piston. If the weight is then removed, the gas expands back to its original volume and the temperature returns to its original value. This is a reversible process. Fig. 3.3 A gas contained in an adiabatic enclosure with a completely frictionless piston. A small mass, dm, is placed on the piston to produce an infinitesimally small pressure change.. It is important not to assume that a reversible process requires an adiabatic enclosure. If the adiabatic enclosure in Fig. 3.3 is replaced with a container with diathermal walls it is also possible to have a reversible change. For example, let the temperature of the surroundings increase by an amount dt energy will flow in through the walls of the container. If the surroundings are then slowly cooled to the original temperature the gas will contract back to its initial volume. This is again a reversible process. Note that a reversible process is an idealization it is not possible to have a completely reversible change in a real system. Nevertheless, as will be shown below, the concept of a reversible process somewhat remarkably allows us to treat many processes that are not reversible of themselves. Quasistatic changes You might ask yourself whether a reversible process must therefore always involve infinitesimally small changes what if we want to reversibly change, for example, the pressure of a system by a large amount? Well, we can make a large change as long as we break it down into a sequence of very small steps. The key feature of this sequence is that the system must remain in equilibrium at all times. If this is the case then the process is referred to as a quasistatic process. If, for example, we decided to push the piston down very rapidly this would most certainly not be a quasistatic process there would be finite temperature and pressure gradients, and turbulence would likely play a role. Hence, a reversible process involving a large change in the properties of the system must proceed via a sequence of quasistatic steps. Reversing this process step-by-step would then produce the

same initial state. Processes which don t involve quasistatic states are irreversible. However, are quasistatic processes always reversible?? When might a quasistatic process not be reversible? NOTES The concept of reversibility will crop up a considerable number of times in the remainder of the module.

3.7 Defining temperature ln (number of microstates) 400 300 200 00 0 ln (number of microstates) vs no. of quanta in block ln(ω 2 ) ln(ω Ω 2 ) ln(ω ) 0 20 40 60 80 00 Number of quanta of energy in block A complete definition of temperature is not possible without considering entropy. Although the concept of temperature as a measure of the mean energy of a system is valid, this does not provide the complete picture as it does not explain why thermal energy transfer occurs from a higher temperature object to a lower temperature object. To derive a formula that relates entropy, energy and temperature, reconsider the two block system described in Section 3.3. Fig. 3.4 Logarithm of number of microstates vs energy of block in a two block system.? What is the value of /dq at thermal equilibrium (where q is the number of quanta of energy in Block )? Thus, at thermal equilibrium: dq + dq 2 = 0 3.3 However, q 2 = 00 q (where q 2 is the number of quanta in Block 2) and so, dq dq 2 2 = 0 3.4 Thus at equilibrium, = 2 3.5 dq dq 2 (NB Don t confuse q (no. of quanta) with Q (heat) and, in the following, don t confuse W (the number of microstates) with work (also W)).

So, at equilibrium, the rate of change of entropy with respect to energy for both blocks is the same. However, at thermal equilibrium, what physical property of the blocks is the same? The temperature! Hence, the rate of change of entropy with respect to energy must somehow be related to temperature. Let s examine the gradient of the graphs of ln (W) vs q shown in Fig. 3.5. At low temperature the gradient of ln (W) vs q (i.e. /dq ) is shown by the line labelled. At high temperature, the gradient is shown by the line labelled 2. However, line 2 represents a smaller gradient. Therefore, although we can define temperature in terms of the rate of change of entropy with respect to energy, we need to realise that a low temperature is associated with a steeper gradient than a high temperature. As the size of one quantum of energy will vary from system to system, it is best to replace q with U, the total internal energy of the system. Thus, our definition of temperature becomes: T = du ln (number of microstates) vs no. of quanta in block ln (number of microstates) 400 300 200 00 0 ln(ω 2 ) ln(ω Ω 2 ) ln(ω ) 2 0 20 40 60 80 00 Number of quanta of energy in block Fig. 3.5 Lines and 2 represent the gradient, /dq at low and high energy (temperature) respectively.

However, this definition of temperature is not complete. It should, in fact, be written as a partial derivative: S = 3.6 T U where variables other than S and U in this case, volume (V) - are treated as constants. As we ll see in Section 4 of the module, if the volume of a gas remains constant then no work is done. We will return to discuss this distinction between heat and work in some detail in Sets 4(a) and 4(b) of the notes but for now note that an input of thermal energy (heat) means that more microstates will become accessible. Hence, the entropy of the system will increase. We can write down a simple expression that relates the change in entropy to the amount of heat transferred: V dq = 3.7 T It is important to realise that Eqn. 3.7 is valid only for very small transfers of energy which are associated with nearly constant T: eqn. 3.7 is valid only for a reversible process. NOTES As written, Eqn. 3.7 is not quite correct it should be written in terms of an inexact differential. However, this subtle, though important, point need not concern us in this module. It will be discussed in some detail in the Statistical Mechanics module next year.

3.8 Reversible and irreversible processes: calculation of entropy There now appears to be an inconsistency between a statement of the 2 nd law given in Section 3.5 and Eqn. 3.7. The statement in question is: the entropy of a thermally isolated system increases in any irreversible process and is unaltered in a reversible process. However, Eqn 3.7 holds only for a reversible process involving a very small amount of heat, dq how do we calculate the entropy change in an irreversible process?! (Remember that a reversible process is an idealization every real process is irreversible). There are some rather subtle thought processes underlying the use of Eqn. 3.7 to calculate entropy changes in irreversible processes which are best illustrated with an example (taken from Finn, p. 75). We want to calculate the entropy change of a beaker of water when it is heated from 293 K (i.e. about room temperature) to 373 K (Fig. 3.6). The beaker is placed on a heat reservoir (i.e. a heat bath which is so large its temperature does not change) and its temperature raised to 373 K (steps 2 and 3 in Fig. 3.6). The beaker is then removed from the reservoir and placed in a jacket with adiabatic walls so there is no longer any heat flow into or out of the system. 2 3 4 293K 293K 373 K 373 K Reservoir at 373K Reservoir at 373 K Fig. 3.6 Raising the temperature of a beaker of water from 293 K to 373 K in one step is an irreversible process. 293. K Reservoir at 293. K T dq=c P T Reservoir at T+ T 293.2 K Reservoir at 293.2 K 373 K Reservoir at 373 K Fig. 3.7 Increasing the temperature of a beaker of water via a series of reversible steps. As this is an irreversible process (due to the large temperature gradients set up as the water is heated) it is not possible to apply Eqn. 3.7 directly. However, the water is in well-defined equilibrium states at the start and end of the process. We imagine a reversible process that moves the state of the water between these two end points. To apply Eqn. 3.7 the irreversible process shown in Fig. 3.7 can be broken down into a series of reversible steps. Starting with the water at T = 293K, the beaker is placed on a reservoir with a temperature of T + T (= 293. K). The temperature difference is so small that the system remains in equilibrium as its temperature changes. A reservoir with T = 293.2 K is then placed next to the beaker and the temperature of the water again increases with the water remaining in equilibrium (see Fig. 3.7).

When the water is at a temperature T and it is heated to a temperature T + T, the heat entering (reversibly) is dq = C P T where C P is the heat capacity at constant pressure. From Eqn. 3.7 (if we now consider infinitesimally small changes in temperature): CPdT = 3.8 T Eqn. 3.8 is the entropy change of the water for each reversible step. To get the total entropy change it is necessary to integrate between the lower and upper limits of temperature (namely, 293 K and 373 K). Thus, the total change in entropy, S is: S = CP 373 293 T I ll leave you to do the integration as an exercise. The reason we can determine changes of entropy in this fashion is related to the concept of a function of state which is covered in Section 4 of the module. dt NOTES

3.9 The Boltzmann distribution revisited Returning to the problem of thermal energy transfer between two blocks, let s now consider one oscillator out of an ensemble of oscillators and ask ourselves the following question: what is the probability that the single oscillator has a certain amount of energy? A helpful way to consider this is to imagine a block comprising 00 atoms (i.e. 300 oscillators) in contact with a single oscillator (Fig. 3.8). Fig. 3.8 A block of 300 oscillators in contact with a single D oscillator. No matter how many quanta of energy we put in the single oscillator, there s only way of distributing them. Therefore, Ω 2 = and W (= Ω Ω 2 ) = Ω 2. A graph of S (= k ln (W)) vs. the total amount of energy in Block (the 00 atom block) is shown in Fig. 3.8 below the total number of quanta is not so important but lets say it s 00 again. Note that S is rising approximately linearly with energy in Block at the high energy side of the graph. This means that S falls with increasing energy in Block 2 (our single oscillator) (because q 2 = 00 q ). At large values of E Block and, thus, low values of Ε (= U E Block ), the graph is approximately linear. (U is the total energy of the system). The graph has slope:. Hence, in de this linear region we can write: S = A de Block E 3.9 where A is an arbitrary constant. Using Eqn. 3.6, Eqn. 3.9 can be rewritten as follows: S E = A 3.0 T However, S = k ln (W) E k ln( W ) = A 3. T Equation 3. can be rewritten (make sure you can do this) as follows: W E = α exp 3.2 kt where α is a constant.

Hopefully equation 3.2 should look familiar. This is the formula for the Boltzmann distribution we first derived in a rather different manner in Section 2.3. Eqn. 3.2 may be used to determine the number of accessible microstates at an energy E above the ground state energy. 3.0 Some abuses of the 2 nd law and the concept of entropy Finally, one of the most misunderstood concepts in popular science writing is that of entropy. Lazy use of terms such as order, disorder, and reversibility has led to very many misconceptions regarding the 2 nd law. For example, very many cranks (and some scientists who should have known better) have dreamt up perpetual motion machines which disobey the 2 nd law these machines have their own special category: perpetual motion machines of the second kind to distinguish them from perpetual motion machines of the first kind which disobey the st law of thermodynamics! In the next section of the module one of the topics with which we ll be concerned will be heat engines and refrigerators hopefully after you study that section of the course you should be able to debunk many of the designs for perpetual motion machines that have been proposed and continue to be proposed on a daily basis by those who refuse to accept the 2 nd and st laws of thermodynamics. (A brief introduction to the nonsense of perpetual motion machines can be found at: http://manor.york.ac.uk/htdocs/perpetual/perpetual.html. There are very many other sites on the web related to this concept!). A key abuse of the 2 nd law is in its application to arrangements of macroscopic objects such as socks in a drawer, books on a desk, cards in a deck, and the contents of a room. An eloquent discussion (by Prof. Frank Lambert) of just how wrong it is to apply the concept of thermodynamic entropy to these situations may be found at: http://jchemed.chem.wisc.edu/journal/issues/999/oct/abs385.html. (Note that links to these sites will be provided on the module website). I will not repeat the arguments of Prof. Lambert in detail here. His take home message is that simply rearranging macroscopic objects does not result in an increase in the number of microstates accessible to those objects hence there is no increase in thermodynamic entropy. That your sock drawer evolves from an ordered, tidy state (perhaps!) to a disordered mess has absolutely nothing to do with entropy. A loose and completely incorrect statement of the 2 nd law is as follows: the entropy of a body always increases. An easy way to address this misconception is to ask how a fridge works. Heat is extracted from a fridge, therefore there must be an entropy decrease somewhere. The problem with the statement as written is that it mixes up net entropy changes with the entropy change of an element of a system. It is the net entropy of a closed system that increases the entropy of individual elements of that system may either increase or decrease. We ll return to a consideration of fridges in Section 4 of the module.

Finally, the most insidious abuse of the 2 nd law is by creationists who use it to apparently debunk the theory of evolution. The argument runs as follows:. Humans and animals are complex, ordered beings 2. Entropy is a measure of disorder 3. The 2 nd law states disorder always increases. 4. Therefore order can t evolve from disorder Darwin must have got things wrong I ll leave you to work out the (gaping) holes in this argument. For more debate on this subject see http://www.talkorigins.org/faqs/thermo.html. NOTES