Search This Blog

Thursday, April 2, 2015

Second law of thermodynamics

From Wikipedia, the free encyclopedia

The second law of thermodynamics states that in a natural thermodynamic process, there is an increase in the sum of the entropies of the participating systems.

The second law is an empirical finding that has been accepted as an axiom of thermodynamic theory.

The law defines the concept of thermodynamic entropy for a thermodynamic system in its own state of internal thermodynamic equilibrium. It considers a process in which that state changes, with increases in entropy due to dissipation of energy and to dispersal of matter and energy.

The law envisages a compound thermodynamic system that initially has interior walls that constrain transfers within it. The law then envisages a process that is initiated by a thermodynamic operation that changes those constraints, and isolates the compound system from its surroundings, except that an externally imposed unchanging force field is allowed to stay subject to the condition that the compound system moves as a whole within that field so that in net, there is no transfer of energy as work between the compound system and the surroundings, and finally, eventually, the system is stationary within that field.

During the process, there may occur chemical reactions, and transfers of matter and of energy. In each adiabatically separated compartment, the temperature becomes spatially homogeneous, even in the presence of the externally imposed unchanging external force field. If, between two adiabatically separated compartments, transfer of energy as work is possible, then it proceeds until the sum of the entropies of the equilibrated compartments is maximum subject to the other constraints. If the externally imposed force field is zero, then the chemical concentrations also become as spatially homogeneous as is allowed by the permeabilities of the interior walls, and by the possibilities of phase separations, which occur so as to maximize the sum of the entropies of the equilibrated phases subject to the other constraints. Such homogeneity and phase separation is characteristic of the state of internal thermodynamic equilibrium of a thermodynamic system.[1][2] If the externally imposed force field is non-zero, then the chemical concentrations spatially redistribute themselves so as to maximize the sum of the equilibrated entropies subject to the other constraints and phase separations.

Statistical thermodynamics, classical or quantum, explains the law.

The second law has been expressed in many ways. Its first formulation is credited to the French scientist Sadi Carnot in 1824 (see Timeline of thermodynamics).

Introduction

The first law of thermodynamics provides the basic definition of thermodynamic energy, also called internal energy, associated with all thermodynamic systems, but unknown in classical mechanics, and states the rule of conservation of energy in nature.[3][4]

The concept of energy in the first law does not, however, account for the observation that natural processes have a preferred direction of progress. The first law is symmetrical with respect to the initial and final states of an evolving system. But the second law asserts that a natural process runs only in one sense, and is not reversible. For example, heat always flows spontaneously from hotter to colder bodies, and never the reverse, unless external work is performed on the system. The key concept for the explanation of this phenomenon through the second law of thermodynamics is the definition of a new physical quantity, the entropy.[5][6]

For mathematical analysis of processes, entropy is introduced as follows. In a fictive reversible process, an infinitesimal increment in the entropy (dS) of a system results from an infinitesimal transfer of heat (δQ) to a closed system divided by the common temperature (T) of the system and the surroundings which supply the heat.[7]
\mathrm dS = \frac{\delta Q}{T} \!
The zeroth law of thermodynamics in its usual short statement allows recognition that two bodies in a relation of thermal equilibrium have the same temperature, especially that a test body has the same temperature as a reference thermometric body.[8] For a body in thermal equilibrium with another, there are indefinitely many empirical temperature scales, in general respectively depending on the properties of a particular reference thermometric body.
The second law allows a distinguished temperature scale, which defines an absolute, thermodynamic temperature, independent of the properties of any particular reference thermometric body.[9][10]

Various statements of the law

The second law of thermodynamics may be expressed in many specific ways,[11] the most prominent classical statements[12] being the statement by Rudolf Clausius (1854), the statement by Lord Kelvin (1851), and the statement in axiomatic thermodynamics by Constantin Carathéodory (1909). These statements cast the law in general physical terms citing the impossibility of certain processes. The Clausius and the Kelvin statements have been shown to be equivalent.[13]

Carnot's principle

The historical origin of the second law of thermodynamics was in Carnot's principle. It refers to a cycle of a Carnot engine, fictively operated in the limiting mode of extreme slowness known as quasi-static, so that the heat and work transfers are between subsystems that are always in their own internal states of thermodynamic equilibrium. The Carnot engine is an idealized device of special interest to engineers who are concerned with the efficiency of heat engines. Carnot's principle was recognized by Carnot at a time when the caloric theory of heat was seriously considered, before the recognition of the first law of thermodynamics, and before the mathematical expression of the concept of entropy. Interpreted in the light of the first law, it is physically equivalent to the second law of thermodynamics, and remains valid today. It states
The efficiency of a quasi-static or reversible Carnot cycle depends only on the temperatures of the two heat reservoirs, and is the same, whatever the working substance. A Carnot engine operated in this way is the most efficient possible heat engine using those two temperatures.[14][15][16][17][18][19][20]
[clarification needed]

Clausius statement

The German scientist Rudolf Clausius laid the foundation for the second law of thermodynamics in 1850 by examining the relation between heat transfer and work.[21] His formulation of the second law, which was published in German in 1854, is known as the Clausius statement:
Heat can never pass from a colder to a warmer body without some other change, connected therewith, occurring at the same time.[22]
The statement by Clausius uses the concept of 'passage of heat'. As is usual in thermodynamic discussions, this means 'net transfer of energy as heat', and does not refer to contributory transfers one way and the other.

Heat cannot spontaneously flow from cold regions to hot regions without external work being performed on the system, which is evident from ordinary experience of refrigeration, for example. In a refrigerator, heat flows from cold to hot, but only when forced by an external agent, the refrigeration system.

Kelvin statement

Lord Kelvin expressed the second law as
It is impossible, by means of inanimate material agency, to derive mechanical effect from any portion of matter by cooling it below the temperature of the coldest of the surrounding objects.[23]

Equivalence of the Clausius and the Kelvin statements


Derive Kelvin Statement from Clausius Statement

Suppose there is an engine violating the Kelvin statement: i.e., one that drains heat and converts it completely into work in a cyclic fashion without any other result. Now pair it with a reversed Carnot engine as shown by the figure. The net and sole effect of this newly created engine consisting of the two engines mentioned is transferring heat

\Delta Q=Q\left(\frac{1}{\eta}-1\right)

from the cooler reservoir to the hotter one, which violates the Clausius statement. Thus a violation of the Kelvin statement implies a violation of the Clausius statement, i.e. the Clausius statement implies the Kelvin statement. We can prove in a similar manner that the Kelvin statement implies the Clausius statement, and hence the two are equivalent.

Planck's proposition

Planck offered the following proposition as derived directly from experience. This is sometimes regarded as his statement of the second law, but he regarded it as a starting point for the derivation of the second law.
It is impossible to construct an engine which will work in a complete cycle, and produce no effect except the raising of a weight and cooling of a heat reservoir.[24][25]

Relation between Kelvin's statement and Planck's proposition

It is almost customary in textbooks to speak of the "Kelvin-Planck statement" of the law, as for example in the text by ter Haar and Wergeland.[26] One text gives a statement that for all the world looks like Planck's proposition, but attributes it to Kelvin without mention of Planck.[27] One monograph quotes Planck's proposition as the "Kelvin-Planck" formulation, the text naming Kelvin as its author, though it correctly cites Planck in its references.[28] The reader may compare the two statements quoted just above here.

Planck's statement

Planck stated the second law as follows.
Every process occurring in nature proceeds in the sense in which the sum of the entropies of all bodies taking part in the process is increased. In the limit, i.e. for reversible processes, the sum of the entropies remains unchanged.[29][30][31]

Principle of Carathéodory

Constantin Carathéodory formulated thermodynamics on a purely mathematical axiomatic foundation. His statement of the second law is known as the Principle of Carathéodory, which may be formulated as follows:[32]
In every neighborhood of any state S of an adiabatically enclosed system there are states inaccessible from S.[33]
With this formulation, he described the concept of adiabatic accessibility for the first time and provided the foundation for a new subfield of classical thermodynamics, often called geometrical thermodynamics. It follows from Carathéodory's principle that quantity of energy quasi-statically transferred as heat is a holonomic process function, in other words, \delta Q=TdS.[34] [clarification needed]

Though it is almost customary in textbooks to say that Carathéodory's principle expresses the second law and to treat it as equivalent to the Clausius or to the Kelvin-Planck statements, such is not the case. To get all the content of the second law, Carathéodory's principle needs to be supplemented by Planck's principle, that isochoric work always increases the internal energy of a closed system that was initially in its own internal thermodynamic equilibrium.[35][36][37][38] [clarification needed]

Planck's Principle

In 1926, Max Planck wrote an important paper on the basics of thermodynamics.[37][39] He indicated the principle
The internal energy of a closed system is increased by an adiabatic process, throughout the duration of which, the volume of the system remains constant.[35][36]
This formulation does not mention heat and does not mention temperature, nor even entropy, and does not necessarily implicitly rely on those concepts, but it implies the content of the second law. A closely related statement is that "Frictional pressure never does positive work."[40] Using a now-obsolete form of words, Planck himself wrote: "The production of heat by friction is irreversible."[41][42]

Not mentioning entropy, this principle of Planck is stated in physical terms. It is very closely related to the Kelvin statement given just above.[43] Nevertheless, this principle of Planck is not actually Planck's preferred statement of the second law, which is quoted above, in a previous sub-section of the present section of this present article, and relies on the concept of entropy.

The link to Kelvin's statement is illustrated by an equivalent statement by Allahverdyan & Nieuwenhuizen, which they attribute to Kelvin: "No work can be extracted from a closed equilibrium system during a cyclic variation of a parameter by an external source."[44][45]

Statement for a system that has a known expression of its internal energy as a function of its extensive state variables

The second law has been shown to be equivalent to the internal energy U being a weakly convex function, when written as a function of extensive properties (mass, volume, entropy, ...).[46][47] [clarification needed]

Gravitational systems

In non-gravitational systems, objects always have positive heat capacity, meaning that the temperature rises with energy. Therefore, when energy flows from a high-temperature object to a low-temperature object, the source temperature is decreased while the sink temperature is increased; hence temperature differences tend to diminish over time.

However, this is not always the case for systems in which the gravitational force is important. The most striking examples are black holes, which – according to theory – have negative heat capacity. The larger the black hole, the more energy it contains, but the lower its temperature. Thus, the supermassive black hole in the center of the Milky Way is supposed to have a temperature of 10−14 K, much lower than the cosmic microwave background temperature of 2.7 K[citation needed], but as it absorbs photons of the cosmic microwave background its mass is increasing so that its low temperature further decreases with time[dubious ].

For this reason, gravitational systems tend towards non-even distribution of mass and energy. The universe in large scale is importantly a gravitational system, and the second law may therefore not apply to it.[dubious ]

Corollaries

Perpetual motion of the second kind

Before the establishment of the Second Law, many people who were interested in inventing a perpetual motion machine had tried to circumvent the restrictions of First Law of Thermodynamics by extracting the massive internal energy of the environment as the power of the machine. Such a machine is called a "perpetual motion machine of the second kind". The second law declared the impossibility of such machines.

Carnot theorem

Carnot's theorem (1824) is a principle that limits the maximum efficiency for any possible engine. The efficiency solely depends on the temperature difference between the hot and cold thermal reservoirs. Carnot's theorem states:
  • All irreversible heat engines between two heat reservoirs are less efficient than a Carnot engine operating between the same reservoirs.
  • All reversible heat engines between two heat reservoirs are equally efficient with a Carnot engine operating between the same reservoirs.
In his ideal model, the heat of caloric converted into work could be reinstated by reversing the motion of the cycle, a concept subsequently known as thermodynamic reversibility. Carnot, however, further postulated that some caloric is lost, not being converted to mechanical work. Hence, no real heat engine could realise the Carnot cycle's reversibility and was condemned to be less efficient.

Though formulated in terms of caloric (see the obsolete caloric theory), rather than entropy, this was an early insight into the second law.

Clausius Inequality

The Clausius Theorem (1854) states that in a cyclic process
\oint \frac{\delta Q}{T} \leq 0.
The equality holds in the reversible case[48] and the '<' is in the irreversible case. The reversible case is used to introduce the state function entropy. This is because in cyclic processes the variation of a state function is zero from state functionality.

Thermodynamic temperature

For an arbitrary heat engine, the efficiency is:
\eta = \frac {A}{q_H} = \frac{q_H-q_C}{q_H} = 1 - \frac{q_C}{q_H} \qquad (1)
where A is the work done per cycle. Thus the efficiency depends only on qC/qH.

Carnot's theorem states that all reversible engines operating between the same heat reservoirs are equally efficient. Thus, any reversible heat engine operating between temperatures T1 and T2 must have the same efficiency, that is to

say, the efficiency is the function of temperatures only: \frac{q_C}{q_H} = f(T_H,T_C)\qquad (2).

In addition, a reversible heat engine operating between temperatures T1 and T3 must have the same efficiency as one consisting of two cycles, one between T1 and another (intermediate) temperature T2, and the second between T2 andT3. This can only be the case if


f(T_1,T_3) = \frac{q_3}{q_1} = \frac{q_2 q_3} {q_1 q_2} = f(T_1,T_2)f(T_2,T_3).
Now consider the case where T_1 is a fixed reference temperature: the temperature of the triple point of water. Then for any T2 and T3,
f(T_2,T_3) = \frac{f(T_1,T_3)}{f(T_1,T_2)} = \frac{273.16 \cdot f(T_1,T_3)}{273.16 \cdot f(T_1,T_2)}.
Therefore if thermodynamic temperature is defined by
T = 273.16 \cdot f(T_1,T) \,
then the function f, viewed as a function of thermodynamic temperature, is simply
f(T_2,T_3) = \frac{T_3}{T_2},
and the reference temperature T1 will have the value 273.16. (Of course any reference temperature and any positive numerical value could be used—the choice here corresponds to the Kelvin scale.)

Entropy

According to the Clausius equality, for a reversible process
\oint \frac{\delta Q}{T}=0
That means the line integral \int_L \frac{\delta Q}{T} is path independent.

So we can define a state function S called entropy, which satisfies
dS = \frac{\delta Q}{T} \!
With this we can only obtain the difference of entropy by integrating the above formula. To obtain the absolute value, we need the Third Law of Thermodynamics, which states that S=0 at absolute zero for perfect crystals.

For any irreversible process, since entropy is a state function, we can always connect the initial and terminal states with an imaginary reversible process and integrating on that path to calculate the difference in entropy.

Now reverse the reversible process and combine it with the said irreversible process. Applying Clausius inequality on this loop,
-\Delta S+\int\frac{\delta Q}{T}=\oint\frac{\delta Q}{T}< 0
Thus,
\Delta S \ge \int \frac{\delta Q}{T} \,\!
where the equality holds if the transformation is reversible.

Notice that if the process is an adiabatic process, then \delta Q=0, so \Delta S\ge 0.

Energy, available useful work

An important and revealing idealized special case is to consider applying the Second Law to the scenario of an isolated system (called the total system or universe), made up of two parts: a sub-system of interest, and the sub-system's surroundings. These surroundings are imagined to be so large that they can be considered as an unlimited heat reservoir at temperature TR and pressure PR — so that no matter how much heat is transferred to (or from) the sub-system, the temperature of the surroundings will remain TR; and no matter how much the volume of the sub-system expands (or contracts), the pressure of the surroundings will remain PR.
Whatever changes to dS and dSR occur in the entropies of the sub-system and the surroundings individually, according to the Second Law the entropy Stot of the isolated total system must not decrease:
 dS_{\mathrm{tot}}= dS + dS_R \ge 0
According to the First Law of Thermodynamics, the change dU in the internal energy of the sub-system is the sum of the heat δq added to the sub-system, less any work δw done by the sub-system, plus any net chemical energy entering the sub-system d ∑μiRNi, so that:
 dU = \delta q - \delta w + d(\sum \mu_{iR}N_i) \,
where μiR are the chemical potentials of chemical species in the external surroundings.

Now the heat leaving the reservoir and entering the sub-system is
 \delta q = T_R (-dS_R) \le T_R dS
where we have first used the definition of entropy in classical thermodynamics (alternatively, in statistical thermodynamics, the relation between entropy change, temperature and absorbed heat can be derived); and then the
Second Law inequality from above.

It therefore follows that any net work δw done by the sub-system must obey
 \delta w \le - dU + T_R dS + \sum \mu_{iR} dN_i \,
It is useful to separate the work δw done by the subsystem into the useful work δwu that can be done by the sub-system, over and beyond the work pR dV done merely by the sub-system expanding against the surrounding external pressure, giving the following relation for the useful work (exergy) that can be done:
 \delta w_u \le -d (U - T_R S + p_R V - \sum \mu_{iR} N_i )\,
It is convenient to define the right-hand-side as the exact derivative of a thermodynamic potential, called the availability or exergy E of the subsystem,
 E = U - T_R S + p_R V - \sum \mu_{iR} N_i
The Second Law therefore implies that for any process which can be considered as divided simply into a subsystem, and an unlimited temperature and pressure reservoir with which it is in contact,
 dE + \delta w_u \le 0 \,
i.e. the change in the subsystem's exergy plus the useful work done by the subsystem (or, the change in the subsystem's exergy less any work, additional to that done by the pressure reservoir, done on the system) must be less than or equal to zero.

In sum, if a proper infinite-reservoir-like reference state is chosen as the system surroundings in the real world, then the Second Law predicts a decrease in E for an irreversible process and no change for a reversible process.
dS_{tot} \ge 0 Is equivalent to  dE + \delta w_u \le 0
This expression together with the associated reference state permits a design engineer working at the macroscopic scale (above the thermodynamic limit) to utilize the Second Law without directly measuring or considering entropy change in a total isolated system. (Also, see process engineer). Those changes have already been considered by the assumption that the system under consideration can reach equilibrium with the reference state without altering the reference state. An efficiency for a process or collection of processes that compares it to the reversible ideal may also be found (See second law efficiency.)

This approach to the Second Law is widely utilized in engineering practice, environmental accounting, systems ecology, and other disciplines.

History

Nicolas Léonard Sadi Carnot in the traditional uniform of a student of the École Polytechnique.

The first theory of the conversion of heat into mechanical work is due to Nicolas Léonard Sadi Carnot in 1824. He was the first to realize correctly that the efficiency of this conversion depends on the difference of temperature between an engine and its environment.

Recognizing the significance of James Prescott Joule's work on the conservation of energy, Rudolf Clausius was the first to formulate the second law during 1850, in this form: heat does not flow spontaneously from cold to hot bodies. While common knowledge now, this was contrary to the caloric theory of heat popular at the time, which considered heat as a fluid. From there he was able to infer the principle of Sadi Carnot and the definition of entropy (1865).

Established during the 19th century, the Kelvin-Planck statement of the Second Law says, "It is impossible for any device that operates on a cycle to receive heat from a single reservoir and produce a net amount of work." This was shown to be equivalent to the statement of Clausius.

The ergodic hypothesis is also important for the Boltzmann approach. It says that, over long periods of time, the time spent in some region of the phase space of microstates with the same energy is proportional to the volume of this region, i.e. that all accessible microstates are equally probable over a long period of time. Equivalently, it says that time average and average over the statistical ensemble are the same.

It has been shown that not only classical systems but also quantum mechanical ones tend to maximize their entropy over time. Thus the second law follows, given initial conditions with low entropy. More precisely, it has been shown that the local von Neumann entropy is at its maximum value with a very high probability.[49] The result is valid for a large class of isolated quantum systems (e.g. a gas in a container). While the full system is pure and therefore does not have any entropy, the entanglement between gas and container gives rise to an increase of the local entropy of the gas. This result is one of the most important achievements of quantum thermodynamics.[dubious ]

Today, much effort in the field is attempting to understand why the initial conditions early in the universe were those of low entropy,[50][51] as this is seen as the origin of the second law (see below).

Informal descriptions

The second law can be stated in various succinct ways, including:
  • It is impossible to produce work in the surroundings using a cyclic process connected to a single heat reservoir (Kelvin, 1851).
  • It is impossible to carry out a cyclic process using an engine connected to two heat reservoirs that will have as its only effect the transfer of a quantity of heat from the low-temperature reservoir to the high-temperature reservoir (Clausius, 1854).
  • If thermodynamic work is to be done at a finite rate, free energy must be expended. (Stoner, 2000)[52]

Mathematical descriptions


Rudolf Clausius

In 1856, the German physicist Rudolf Clausius stated what he called the "second fundamental theorem in the mechanical theory of heat" in the following form:[53]
\int \frac{\delta Q}{T} = -N
where Q is heat, T is temperature and N is the "equivalence-value" of all uncompensated transformations involved in a cyclical process. Later, in 1865, Clausius would come to define "equivalence-value" as entropy. On the heels of this definition, that same year, the most famous version of the second law was read in a presentation at the Philosophical Society of Zurich on April 24, in which, in the end of his presentation, Clausius concludes:
The entropy of the universe tends to a maximum.
This statement is the best-known phrasing of the second law. Because of the looseness of its language, e.g. universe, as well as lack of specific conditions, e.g. open, closed, or isolated, many people take this simple statement to mean that the second law of thermodynamics applies virtually to every subject imaginable. This, of course, is not true; this statement is only a simplified version of a more extended and precise description.

In terms of time variation, the mathematical statement of the second law for an isolated system undergoing an arbitrary transformation is:
\frac{dS}{dt} \ge 0
where
S is the entropy of the system and
t is time.
The equality sign holds in the case that only reversible processes take place inside the system. If irreversible processes take place (which is the case in real systems in operation) the >-sign holds. An alternative way of formulating of the second law for isolated systems is:
\frac{dS}{dt} = \dot S_{i} with  \dot S_{i} \ge 0
with  \dot S_{i} the sum of the rate of entropy production by all processes inside the system. The advantage of this formulation is that it shows the effect of the entropy production. The rate of entropy production is a very important concept since it determines (limits) the efficiency of thermal machines. Multiplied with ambient temperature T_{a} it gives the so-called dissipated energy  P_{diss}=T_{a}\dot S_{i}.

The expression of the second law for closed systems (so, allowing heat exchange and moving boundaries, but not exchange of matter) is:
\frac{dS}{dt} = \frac{\dot Q}{T}+\dot S_{i} with  \dot S_{i} \ge 0
Here
\dot Q is the heat flow into the system
T is the temperature at the point where the heat enters the system.
If heat is supplied to the system at several places we have to take the algebraic sum of the corresponding terms.

For open systems (also allowing exchange of matter):
\frac{dS}{dt} = \frac{\dot Q}{T}+\dot S+\dot S_{i} with  \dot S_{i} \ge 0
Here \dot S is the flow of entropy into the system associated with the flow of matter entering the system. It should not be confused with the time derivative of the entropy. If matter is supplied at several places we have to take the algebraic sum of these contributions.

Statistical mechanics gives an explanation for the second law by postulating that a material is composed of atoms and molecules which are in constant motion. A particular set of positions and velocities for each particle in the system is called a microstate of the system and because of the constant motion, the system is constantly changing its microstate. Statistical mechanics postulates that, in equilibrium, each microstate that the system might be in is equally likely to occur, and when this assumption is made, it leads directly to the conclusion that the second law must hold in a statistical sense. That is, the second law will hold on average, with a statistical variation on the order of 1/√N where N is the number of particles in the system. For everyday (macroscopic) situations, the probability that the second law will be violated is practically zero. However, for systems with a small number of particles, thermodynamic parameters, including the entropy, may show significant statistical deviations from that predicted by the second law. Classical thermodynamic theory does not deal with these statistical variations.

Derivation from statistical mechanics

Due to Loschmidt's paradox, derivations of the Second Law have to make an assumption regarding the past, namely that the system is uncorrelated at some time in the past; this allows for simple probabilistic treatment. This assumption is usually thought as a boundary condition, and thus the second Law is ultimately a consequence of the initial conditions somewhere in the past, probably at the beginning of the universe (the Big Bang), though other scenarios have also been suggested.[54][55][56]
Given these assumptions, in statistical mechanics, the Second Law is not a postulate, rather it is a consequence of the fundamental postulate, also known as the equal prior probability postulate, so long as one is clear that simple probability arguments are applied only to the future, while for the past there are auxiliary sources of information which tell us that it was low entropy.[citation needed] The first part of the second law, which states that the entropy of a thermally isolated system can only increase, is a trivial consequence of the equal prior probability postulate, if we restrict the notion of the entropy to systems in thermal equilibrium. The entropy of an isolated system in thermal equilibrium containing an amount of energy of E is:
S = k_{\mathrm B} \ln\left[\Omega\left(E\right)\right]\,
where \Omega\left(E\right) is the number of quantum states in a small interval between E and E +\delta E. Here \delta E is a macroscopically small energy interval that is kept fixed. Strictly speaking this means that the entropy depends on the choice of \delta E. However, in the thermodynamic limit (i.e. in the limit of infinitely large system size), the specific entropy (entropy per unit volume or per unit mass) does not depend on \delta E.

Suppose we have an isolated system whose macroscopic state is specified by a number of variables. These macroscopic variables can, e.g., refer to the total volume, the positions of pistons in the system, etc. Then \Omega will depend on the values of these variables. If a variable is not fixed, (e.g. we do not clamp a piston in a certain position), then because all the accessible states are equally likely in equilibrium, the free variable in equilibrium will be such that \Omega is maximized as that is the most probable situation in equilibrium.

If the variable was initially fixed to some value then upon release and when the new equilibrium has been reached, the fact the variable will adjust itself so that \Omega is maximized, implies that the entropy will have increased or it will have stayed the same (if the value at which the variable was fixed happened to be the equilibrium value). Suppose we start from an equilibrium situation and we suddenly remove a constraint on a variable. Then right after we do this, there are a number \Omega of accessible microstates, but equilibrium has not yet been reached, so the actual probabilities of the system being in some accessible state are not yet equal to the prior probability of 1/\Omega. We have already seen that in the final equilibrium state, the entropy will have increased or have stayed the same relative to the previous equilibrium state. Boltzmann's H-theorem, however, proves that the quantity H increases monotonically as a function of time during the intermediate out of equilibrium state.

Derivation of the entropy change for reversible processes

The second part of the Second Law states that the entropy change of a system undergoing a reversible process is given by:
dS =\frac{\delta Q}{T}
where the temperature is defined as:
\frac{1}{k_{\mathrm B} T}\equiv\beta\equiv\frac{d\ln\left[\Omega\left(E\right)\right]}{dE}
See here for the justification for this definition. Suppose that the system has some external parameter, x, that can be changed. In general, the energy eigenstates of the system will depend on x. According to the adiabatic theorem of quantum mechanics, in the limit of an infinitely slow change of the system's Hamiltonian, the system will stay in the same energy eigenstate and thus change its energy according to the change in energy of the energy eigenstate it is in.

The generalized force, X, corresponding to the external variable x is defined such that X dx is the work performed by the system if x is increased by an amount dx. E.g., if x is the volume, then X is the pressure. The generalized force for a system known to be in energy eigenstate E_{r} is given by:
X = -\frac{dE_{r}}{dx}
Since the system can be in any energy eigenstate within an interval of \delta E, we define the generalized force for the system as the expectation value of the above expression:
X = -\left\langle\frac{dE_{r}}{dx}\right\rangle\,
To evaluate the average, we partition the \Omega\left(E\right) energy eigenstates by counting how many of them have a value for \frac{dE_{r}}{dx} within a range between Y and Y + \delta Y. Calling this number \Omega_{Y}\left(E\right), we have:
\Omega\left(E\right)=\sum_{Y}\Omega_{Y}\left(E\right)\,
The average defining the generalized force can now be written:
X = -\frac{1}{\Omega\left(E\right)}\sum_{Y} Y\Omega_{Y}\left(E\right)\,
We can relate this to the derivative of the entropy w.r.t. x at constant energy E as follows. Suppose we change x to x + dx. Then \Omega\left(E\right) will change because the energy eigenstates depend on x, causing energy eigenstates to move into or out of the range between E and E+\delta E. Let's focus again on the energy eigenstates for which \frac{dE_{r}}{dx} lies within the range between Y and Y + \delta Y. Since these energy eigenstates increase in energy by Y dx, all such energy eigenstates that are in the interval ranging from E – Y dx to E move from below E to above E. There are
N_{Y}\left(E\right)=\frac{\Omega_{Y}\left(E\right)}{\delta E} Y dx\,
such energy eigenstates. If Y dx\leq\delta E, all these energy eigenstates will move into the range between E and E+\delta E and contribute to an increase in \Omega. The number of energy eigenstates that move from below E+\delta E to above E+\delta E is, of course, given by N_{Y}\left(E+\delta E\right). The difference
N_{Y}\left(E\right) - N_{Y}\left(E+\delta E\right)\,
is thus the net contribution to the increase in \Omega. Note that if Y dx is larger than \delta E there will be the energy eigenstates that move from below E to above E+\delta E. They are counted in both N_{Y}\left(E\right) and N_{Y}\left(E+\delta E\right), therefore the above expression is also valid in that case.

Expressing the above expression as a derivative w.r.t. E and summing over Y yields the expression:
\left(\frac{\partial\Omega}{\partial x}\right)_{E} = -\sum_{Y}Y\left(\frac{\partial\Omega_{Y}}{\partial E}\right)_{x}= \left(\frac{\partial\left(\Omega X\right)}{\partial E}\right)_{x}\,
The logarithmic derivative of \Omega w.r.t. x is thus given by:
\left(\frac{\partial\ln\left(\Omega\right)}{\partial x}\right)_{E} = \beta X +\left(\frac{\partial X}{\partial E}\right)_{x}\,
The first term is intensive, i.e. it does not scale with system size. In contrast, the last term scales as the inverse system size and will thus vanishes in the thermodynamic limit. We have thus found that:
\left(\frac{\partial S}{\partial x}\right)_{E} = \frac{X}{T}\,
Combining this with
\left(\frac{\partial S}{\partial E}\right)_{x} = \frac{1}{T}\,
Gives:
dS = \left(\frac{\partial S}{\partial E}\right)_{x}dE+\left(\frac{\partial S}{\partial x}\right)_{E}dx = \frac{dE}{T} + \frac{X}{T} dx=\frac{\delta Q}{T}\,

Derivation for systems described by the canonical ensemble

If a system is in thermal contact with a heat bath at some temperature T then, in equilibrium, the probability distribution over the energy eigenvalues are given by the canonical ensemble:
P_{j}=\frac{\exp\left(-\frac{E_{j}}{k_{\mathrm B} T}\right)}{Z}
Here Z is a factor that normalizes the sum of all the probabilities to 1, this function is known as the partition function. We now consider an infinitesimal reversible change in the temperature and in the external parameters on which the energy levels depend. It follows from the general formula for the entropy:
S = -k_{\mathrm B}\sum_{j}P_{j}\ln\left(P_{j}\right)
that
dS = -k_{\mathrm B}\sum_{j}\ln\left(P_{j}\right)dP_{j}
Inserting the formula for P_{j} for the canonical ensemble in here gives:
dS = \frac{1}{T}\sum_{j}E_{j}dP_{j}=\frac{1}{T}\sum_{j}d\left(E_{j}P_{j}\right) - \frac{1}{T}\sum_{j}P_{j}dE_{j}= \frac{dE + \delta W}{T}=\frac{\delta Q}{T}

General derivation from unitarity of quantum mechanics

The time development operator in quantum theory is unitary, because the Hamiltonian is hermitian. Consequently, the transition probability matrix is doubly stochastic, which implies the Second Law of Thermodynamics.[57][58]
This derivation is quite general, based on the Shannon entropy, and does not require any assumptions beyond unitarity, which is universally accepted. It is a consequence of the irreversibility or singular nature of the general transition matrix.

Non-equilibrium states

The theory of classical or equilibrium thermodynamics is idealized. A main postulate or assumption, often not even explicitly stated, is the existence of systems in their own internal states of thermodynamic equilibrium. In general, a region of space containing a physical system at a given time, that may be found in nature, is not in thermodynamic equilibrium, read in the most stringent terms. In looser terms, nothing in the entire universe is or has ever been truly in exact thermodynamic equilibrium.[59][60]
For purposes of physical analysis, it is often enough convenient to make an assumption of thermodynamic equilibrium. Such an assumption may rely on trial and error for its justification. If the assumption is justified, it can often be very valuable and useful because it makes available the theory of thermodynamics. Elements of the equilibrium assumption are that a system is observed to be unchanging over an indefinitely long time, and that there are so many particles in a system, that its particulate nature can be entirely ignored. Under such an equilibrium assumption, in general, there are no macroscopically detectable fluctuations. There is an exception, the case of critical states, which exhibit to the naked eye the phenomenon of critical opalescence. For laboratory studies of critical states, exceptionally long observation times are needed.

In all cases, the assumption of thermodynamic equilibrium, once made, implies as a consequence that no putative candidate "fluctuation" alters the entropy of the system.

It can easily happen that a physical system exhibits internal macroscopic changes that are fast enough to invalidate the assumption of the constancy of the entropy. Or that a physical system has so few particles that the particulate nature is manifest in observable fluctuations. Then the assumption of thermodynamic equilibrium is to be abandoned. There is no unqualified general definition of entropy for non-equilibrium states.[61]

Non-equilibrium thermodynamics is then appropriate. There are intermediate cases, in which the assumption of local thermodynamic equilibrium is a very good approximation,[62][63][64][65] but strictly speaking it is still an approximation, not theoretically ideal. For non-equilibrium situations in general, it may be useful to consider statistical mechanical definitions of quantities that may be conveniently called 'entropy'. These indeed belong to statistical mechanics, not to macroscopic thermodynamics.

The physics of macroscopically observable fluctuations is beyond the scope of this article.

Arrow of time

The second law of thermodynamics is a physical law that is not symmetric to reversal of the time direction.The second law has been proposed to supply an explanation of the difference between moving forward and backwards in time, such as why the cause precedes the effect (the causal arrow of time).[66]

Controversies

Maxwell's demon

James Clerk Maxwell

James Clerk Maxwell imagined one container divided into two parts, A and B. Both parts are filled with the same gas at equal temperatures and placed next to each other. Observing the molecules on both sides, an imaginary demon guards a trapdoor between the two parts. When a faster-than-average molecule from A flies towards the trapdoor, the demon opens it, and the molecule will fly from A to B. The average speed of the molecules in B will have increased while in A they will have slowed down on average. Since average molecular speed corresponds to temperature, the temperature decreases in A and increases in B, contrary to the second law of thermodynamics.

One of the most famous responses to this question was suggested in 1929 by Leó Szilárd and later by Léon Brillouin. Szilárd pointed out that a real-life Maxwell's demon would need to have some means of measuring molecular speed, and that the act of acquiring information would require an expenditure of energy.

Maxwell's demon repeatedly alters the permeability of the wall between A and B. It is therefore performing thermodynamic operations, not just presiding over natural processes.

Loschmidt's paradox

Loschmidt's paradox, also known as the reversibility paradox, is the objection that it should not be possible to deduce an irreversible process from time-symmetric dynamics. This puts the time reversal symmetry of nearly all known low-level fundamental physical processes at odds with any attempt to infer from them the second law of thermodynamics which describes the behavior of macroscopic systems. Both of these are well-accepted principles in physics, with sound observational and theoretical support, yet they seem to be in conflict; hence the paradox.
One proposed resolution of this paradox is as follows. The Loschmidt scenario refers to a strictly isolated system or to a strictly adiabatically isolated system. Heat and matter transfers are not allowed. The Loschmidt reversal times are fantastically long, far longer than any laboratory isolation of the required degree of perfection could be maintained in practice. In this sense, the Loschmidt scenario will never be subjected to empirical testing. Also in this sense, the second law, stated for an isolated system, will never be subjected to empirical testing. A system, supposedly perfectly isolated, in strictly perfect thermodynamic equilibrium, can be observed only once in its entire life, because the observation must break the isolation. Two observations would be needed to check empirically for a change of state, one initial and one final. When transfer of heat or matter are permitted, the requirements of perfection are not so tight. In practical laboratory reality, therefore, the second law can be tested only for systems with transfer of heat or matter, and not for isolated systems.

Due to this paradox, derivations of the second law have to make an assumption regarding the past, namely that the system is uncorrelated at some time in the past or, equivalently, that the entropy in the past was lower than in the future. This assumption is usually thought as a boundary condition, and thus the second Law is ultimately derived from the initial conditions of the Big Bang.[54][67]

Poincaré recurrence theorem

The Poincaré recurrence theorem states that certain systems will, after a sufficiently long time, return to a state very close to the initial state. The Poincaré recurrence time is the length of time elapsed until the recurrence, which is of the order of \sim \exp\left(S/k\right).[68] The result applies to physical systems in which energy is conserved. The Recurrence theorem apparently contradicts the Second law of thermodynamics, which says that large dynamical systems evolve irreversibly towards the state with higher entropy, so that if one starts with a low-entropy state, the system will never return to it. There are many possible ways to resolve this paradox, but none of them is universally accepted.[citation needed] The most reasonable argument is that for typical thermodynamical systems the recurrence time is so large (many many times longer than the lifetime of the universe) that, for all practical purposes, one cannot observe the recurrence.

Quotations

The law that entropy always increases holds, I think, the supreme position among the laws of Nature. If someone points out to you that your pet theory of the universe is in disagreement with Maxwell's equations — then so much the worse for Maxwell's equations. If it is found to be contradicted by observation — well, these experimentalists do bungle things sometimes. But if your theory is found to be against the second law of thermodynamics I can give you no hope; there is nothing for it but to collapse in deepest humiliation.
—Sir Arthur Stanley Eddington, The Nature of the Physical World (1927)
There have been nearly as many formulations of the second law as there have been discussions of it.
—Philosopher / Physicist P.W. Bridgman, (1941)
Clausius is the author of the sibyllic utterance, "The energy of the universe is constant; the entropy of the universe tends to a maximum." The objectives of continuum thermomechanics stop far short of explaining the "universe", but within that theory we may easily derive an explicit statement in some ways reminiscent of Clausius, but referring only to a modest object: an isolated body of finite size.
Truesdell, C., Muncaster, R.G. (1980). Fundamentals of Maxwell's Kinetic Theory of a Simple Monatomic Gas, Treated as a Branch of Rational Mechanics, Academic Press, New York, ISBN0-12-701350-4, p.17.

In This City It’s Now Illegal To Discriminate Against Atheists

 


Atheist Megachurches
CREDIT: AP

On Tuesday, the city of Madison, Wisconsin announced that it is now against the law to discriminate against atheists, making it the first city in the country to grant explicit legal protection to people who do not believe in a god.

According to Hemant Mehta of the Friendly Atheist blog, last night the Madison city council voted unanimously to add atheists to a list of protected groups in the city’s equal opportunity ordinance, an anti-discrimination law. The move, which inserts the phrase “religion or nonreligion” into the legal code, prevents atheists from being denied equal opportunity in employment, housing, and public accommodations.

“This is important because I believe it is only fair that if we protect religion, in all its varieties, we should also protect non-religion from discrimination,” Anita Weier, an Alderwoman in Madison and sponsor of the ordinance, told local news affiliate Channel 3000.

The ordinance also outlaws discrimination based on a number of other factors such as sex, race, citizenship status, arrest record, sexual orientation, gender identity, or anyone who declines to disclose their social security number, among many others. Reportedly, no one at the council meeting voiced disagreement with the proposal to include atheists.

The new law is part of a growing movement to claim formal protection for atheists, who often face explicit or implicit discrimination for their non-belief. Although the U.S. Constitution expressly prohibits submitting candidates for office to a religious test, people who do not believe in God are currently legally barred from holding office in seven states: North Carolina, Arkansas, Maryland, Mississippi, South Carolina, Tennessee, and Texas. These statutes are, of course, unlikely to hold up in court, but atheists also face substantial hurdles at the ballot box, as a 2012 Gallup poll found that Americans are more likely to vote for Mormons, Muslims, or gay people than atheist candidates.

Atheists have scored a few legal wins over the past few years, such as a federal district court in Oregon declaring in 2014 that Secular Humanism is a “religion” for the purposes of legal protection. In addition, last year advocates from across the country successfully pressured an Alaska town into guaranteeing atheists the right to deliver invocations at city council meetings, and several nonbeliever groups continue to push for the inclusion of a Humanist chaplain in the U.S. military.
Update
This post was updated to reflect that while the initial proposal for the legislation wanted to insert the phrase “religion or atheism,” the final legislation that passed yesterday reads “religion and nonreligion,” a broader term that includes atheists but also others who might be excluded from the atheist category.

A robot prepared for self-awareness


Original link:  http://phys.org/news/2015-04-robot-self-awareness.html


A robot prepared for self-awareness
The walking robot Hector took his first steps at the end of 2014.
Credit: CITEC/Bielefeld University
A year ago, researchers at Bielefeld University showed that their software endowed the walking robot Hector with a simple form of consciousness. Their new research goes one step forward: they have now developed a software architecture that could enable Hector to see himself as others see him. "With this, he would have reflexive consciousness," explains Dr. Holk Cruse, professor at the Cluster of Excellence Cognitive Interaction Technology (CITEC) at Bielefeld University. The architecture is based on artificial neural networks. Together with colleague Dr. Malte Schilling, Prof. Dr. Cruse published this new study in the online collection Open MIND, a volume from the Mind-Group, which is a group of philosophers and other scientists studying the mind, consciousness, and cognition.

Both biologists are involved in further developing and enhancing Hector's . The robot is modelled after a stick insect. How Hector walks and deals with obstacles in its path were first demonstrated at the end of 2014. Next, Hector's extended software will now be tested using a computer simulation. "What works in the computer simulation must then, in a second phase, be transferred over to the robot and tested on it," explains Cruse. Drs. Schilling and Cruse are investigating to what extent various higher level , for example aspects of consciousness, may develop in Hector with this software – even though these traits were not specifically built in to the robot beforehand. The researchers speak of "emergent" abilities, that is, capabilities that suddenly appear or emerge.

Until now, Hector has been a reactive system. It reacts to stimuli in its surroundings. Thanks to the software program "Walknet," Hector can walk with an insect-like gait, and another program called "Navinet" may enable the robot to find a path to a distant target. Both researchers have also developed the software expansion programme "reaCog." This software is activated in instances when both of the other programmes are unable to solve a given problem. This new expanded software enables the robot to simulate "imagined behaviour" that may solve the problem: first, it looks for new solutions and evaluates whether this action makes sense, instead of just automatically completing a pre-determined operation. Being able to perform imagined actions is a central characteristic of a simple form of consciousness.

In their previous research, both CITEC researchers had already determined that Hector's control system could adopt a number of higher-level mental states. "Intentions, for instance, can be found in the system," explains Malte Schilling. These "inner mental states," such as intentions, make goal-directed behaviour possible, which for example may direct the robot to a certain location (like a charging station). The researchers have also identified how properties of emotions may show up in the system. "Emotions can be read from behaviour. For example, a person who is happy takes more risks and makes decisions faster than someone who is anxious," says Holk Cruse. This behaviour could also be implemented in the control model reaCog: "Depending on its inner mental state, the system may adopt quick, but risky solutions, and at other times, it may take its time to search for a safer solution."

To examine which forms of consciousness are present in Hector, the researchers rely on psychological and neurobiological definitions in particular. As Holk Cruse explains, "A human possesses reflexive consciousness when he not only can perceive what he experiences, but also has the ability to experience that he is experiencing something. Reflexive consciousness thus exists if a human or a technical system can see itself "from outside of itself," so to speak."

In their new research, Cruse and Schilling show a way in which reflexive consciousness could emerge. "With the new software, Hector could observe its inner mental state – to a certain extent, its moods – and direct its actions using this information," says Malte Schilling. "What makes this unique, however, is that with our software expansion, the basic faculties are prepared so that Hector may also be able to assess the mental state of others. It may be able to sense other people's intentions or expectations and act accordingly." Dr. Cruse explains further, "the may then be able to "think": what does this subject expect from me? And then it can orient its actions accordingly."

Cruse and Schilling's study is part of the online publication "Open MIND." This approximately 2,000-page collection marks the 10-year anniversary of the MIND-Group and contains 39 original articles from scientists in the fields of philosophy, psychology, and neurological research. Dr. Thomas Metzinger of the University of Mainz is the initiator and co-editor of the volume. The collection can be accessed online for free at www.open-mind.net and is also available in print.

More information: Holk Cruse, Malte Schilling: "Mental States as Emergent Properties. From Walking to Consciousness." In: Thomas Metzinger, Jennifer Windt (eds.): Open Mind. MIND Group, 335–373, open-mind.net/papers/mental-states-as-emergent-properties-from-walking-to-consciousness, published on 20 January 2015.

Provided by Bielefeld University

The Next Bubble

The government's working hard to inflate another housing bubble


When the last housing bubble burst, politicians blamed "greedy banks." They said mortgage companies lent money recklessly, making loans to people with dubious credit, for down payments as low as three percent.

"It will work out," said the optimistic bankers. Regulators didn't disagree. Everyone said, "Home prices will keep going up." And home prices did—until they didn't.

The bubble popped in 2007. Lots of people were hurt, and politicians took more of your tax money to bail out Fannie Mae and Freddie Mac along with reckless banks. They also gave the Federal Housing Administration a $2 billion bailout.

Then the politicians said, "We'll fix this so it doesn't happen again." Congress passed Dodd-Frank and a thousand new regulations. The complex rules slowed lending, all right. It's one reason this post-recession recovery has been abnormally slow.

But—April Fools'!—the new rules didn't solve the problem of reckless lending, and it's happening again.

Because our government subsidizes home purchases, recklessness is invited. Somehow, Americans buy cars, clothing, computers, etc. without government guarantees, but politicians think housing is different.

Both parties support the subsidies.

The left wants government to help struggling families, and the right thinks home ownership sends a wholesome cultural message. Both parties have cozy connections to home-builders and lenders.

At the time of the housing crash, most high-risk loans were guaranteed by the government. Those banks wouldn't have been as reckless if they had their own money on the line.

But they knew they could grant a mortgage to most anyone and the FHA would back it or government-sponsored companies Fannie Mae and Freddie Mac would buy it. That fueled the frenzy of lending.

After the bubble popped, I assumed the political class would learn a lesson, but they haven't. Today, even more American mortgages are guaranteed by government. More than 90 percent of new loans are backed by taxpayers. After the crash, Fannie and Freddie did raise their minimum down payment—to a measly five percent—but a few months ago, they lowered it again to three percent!

Are they crazy? A sensible congressman, Rep. Jeb Hensarling (R-Tex.), tried to get an answer from the administration's new mortgage regulator, asking in a hearing, "All things being equal, is a three percent down riskier to the taxpayer than a ten percent down loan?"

A pretty basic question—but one that director Mel Watt still dodged, responding, "Mr. Chairman, that is generally true. But when you pair the down payment with compensating factors ... look at other considerations ... you can ensure that a three percent loan is just as safe."

What? That's nonsense. This is what happens when pandering politicians get to dispense your money. Watt is among the worst. When he was a congressman, he pushed for mortgage subsidies for welfare recipients who made down payments as low as $1,000.

Edward Pinto, who studies housing risk for the American Enterprise Institute, says policies like this put us on the way to another bubble: "The government is once again ... saying, let's loosen credit, give loans to people that potentially can't afford them, and everything will be fine because house prices will go up."

On my show, former FHA commissioner David Stevens, who did improve lending standards a bit after the crash (before Watt and his cronies weakened them), responded that this time the government has new regulations that will prevent things falling apart: "I think in the effort, post-recession, to make sure we never go down this path again, we have created more rules than ever existed in the history of this country."

But more rules aren't a solution. Government's regulators didn't foresee the problems last time. Fannie and Freddie got a clean bill of health right up until the collapse.

The solution is less government involvement. Canada doesn't have a Fannie, Freddie or FHA. Canada didn't have the trauma of a housing bubble. In Canada, lenders and homeowners risk their own  money.

Does that mean Canadians cannot afford homes? No! Without all that government help, Canada's homeownership rate is higher than ours.

John Stossel (read his Reason archive) is the host of Stossel, which airs Thursdays on the FOX Business Network at 9 pm ET and is rebroadcast on Saturdays and Sundays at 9pm & midnight ET. Go here for more info.

Two-state solution

From Wikipedia, the free encyclopedia https://en.wikipedia.org/wiki/Two-state_solution A peace movement po...