Download the constancy of for an ideal gas undergoing an adiabatic

Survey
yes no Was this document useful for you?
   Thank you for your participation!

* Your assessment is very important for improving the workof artificial intelligence, which forms the content of this project

Document related concepts

Scalar field theory wikipedia , lookup

Atomic theory wikipedia , lookup

History of quantum field theory wikipedia , lookup

Hydrogen atom wikipedia , lookup

Bohr–Einstein debates wikipedia , lookup

Path integral formulation wikipedia , lookup

Interpretations of quantum mechanics wikipedia , lookup

Quantum teleportation wikipedia , lookup

EPR paradox wikipedia , lookup

Symmetry in quantum mechanics wikipedia , lookup

Quantum state wikipedia , lookup

Relativistic quantum mechanics wikipedia , lookup

Wave–particle duality wikipedia , lookup

Hidden variable theory wikipedia , lookup

Canonical quantization wikipedia , lookup

Particle in a box wikipedia , lookup

Theoretical and experimental justification for the Schrödinger equation wikipedia , lookup

Matter wave wikipedia , lookup

T-symmetry wikipedia , lookup

Transcript
THE CASSICAL ADIABATIC CONSTANCY OF PV γ FOR AN IDEAL GAS, TURNS
OUT TO BE A QUANTUM MECHANICAL OCCURRENCE
Tolga Yarman, Okan University, Akfirat, Istanbul, Turkey ([email protected]) &
Savronik, Organize Sanayii Bölgesi, Eskisehir, Turkey,
Alexander L Kholmetskii, Belarus State University, Minsk, Belarus ([email protected]),
Önder Korfali, Galatasaray University, Ortaköy, Istanbul, Turkey
ABSTRACT
In this paper we find a connection between the macroscopic classical laws of gases and the quantum mechanical
description of non-interacting particles confined in a box, in fact constituting an ideal gas. In such a gas, the
motion of each individual molecule can be considered independently from all other molecules, and thus the macroscopic parameters of ideal gas, like pressure P and temperature T, can be introduced as a result of simple averaging over all individual motions of molecules. It is shown that for an ideal gas enclosed in a macroscopic cubic
box of volume V, the constant, in the classical law of adiabatic expansion expression, i.e. PV 5 / 3  Constant ,
can be derived, based on quantum mechanics. Physical implications of the result we disclose are discussed. In
any case, our finding proves, seemingly, a macroscopic manifestation of a quantum mechanical behavior, and
this in relation to classical thermodynamics.
1. INTRODUCTION
Time to time, most of us, no doubt, just like many scientists of the 20th century, were puzzled
with the question of finding a bridge between the Boltzmann Constant k and the Planck
Constant h. We will see below that this is actually a vain effort. Nevertheless, de Broglie already in his doctorate thesis has brilliantly applied his relationship (associating a wave length
with the momentum of a moving particle) to the statistical equilibrium of gases [1], but did not
advance his idea, to see whether one can, along such a line, obtain anything related to the law
of gases, established long ago, in 1650.
The Boyle-Mariotte law of ideal gas is given, as usual, by
(1)
PV  n RT  kNT
with the following designations: P is the pressure of the gas, V the volume of the gas, T the
temperature of the gas, n the number of moles the gas is made of, N the number of molecules
making the gas, R the gas constant, and k the Boltzmann Constant.
The value of R is measured to be
R = 8.31 Joules / º K .
(2)
The Boltzmann Constant is then given by
(3)
k  R / NA ,
where N A is the Avagadro number; note that n is, N / NA .
The Kinetic Theory of Gases allows us to derive the same casing as that of Eq.(1) via
considering the momentum change of molecules when bouncing back from a wall of the container [2]. Assuming for simplicity a cubic geometry, one obtains*
*
For the sake of completeness, let us recall the classical derivation of Eq.(4). The force f x exerted by the molecule of mass m and velocity v, delineating vx as its x-component, on the wall  to x, is given by Newton’s
second law, i.e. f x  p x / t , where p x  2mv x is the algebraic increase in the momentum, whilst the
molecule bounces back from the wall, and t  2 L / vx , L being the size of the container along the x-direction.
1
PV 
N
2
mv 2  N E ,
3
3
(4)
E  mv 2 / 2 , being the average translational energy of molecules of mass m.
The comparison of this relationship with Eq.(1) yields
3
E  kT .
(5)
2
Furthermore, Eq.(4), given the way it is framed (cf. the footnote, we have just provided), can well be written for the pressure p, that would be built in a volume V, containing just
one molecule of translational energy E :
pV  kT 
2
E.
3
(6)
Could this equation be a basis to build a bridge between the law of gases (mainly
characterized by the Boltzmann Constant), and quantum mechanics (which will evidently involve the energy quantity E )? Here though, while the equality pV  kT points to the law of
gases, the next equality kT  (2 / 3) E is no more than a definition of the temperature, in terms
of the average translational energy of the molecules E , once k is defined via Eq.(3). So
kT  (2 / 3) E is not to provide us with any relationship between k and h . (It is just a definition
of T, once k is known.)
Accordingly, Eq.(6) does not bring us anywhere in finding a bridge between the “law
of gases” and “quantum mechanics”.
In other terms, E is to be expected to involve the Planck Constant, yet this, does not
provide us, with a relationship between h and k, if we based ourselves on Eq.(6), for such an
equality would yield merely a relationship between h and kT. In effect, the introduction of the
temperature concept next to the concept of energy is not only redundant, but also somehow
dichotemic, thus misleading. Hence, based on Eq.(6) we are bound to fail to establish a relationship between macroscopic properties of an ideal gas and the quantum mechanical description of its molecules.
Thereby we find out that, when we propose to draw a line between the law of gases
and quantum mechanics, we should not really look for a relationship between h and k. Any
such effort will be dissolved through a plain definition of the temperature, in terms of the average translational energy of the molecules, and nothing beyond. How ever, we can still go
ahead to check whether the phenomenological laws of gases are well matched to quantum
mechanics, if we could explore those laws of gases, which do not involve the constants R or k.
That is the key point of our approach.
Thus, f x becomes f x  mv x2 / L . We can suppose that we deal with an average molecule, and all molecules behave as this average molecule. Hence, summing over N molecules, the gas is made of, we get the total force
Fx  N
mvx2 N mv 2 ,

L
3 L
where we have the mean square velocities; recall that at the equilibrium the mean square velocities, for all directions, point to the same quantity. The pressure P exerted by N molecules on the wall of concern, is thence
N mv 2 ,
3 L3
which is Eq.(4), along with V=L3. Note that the foregoing derivation is well based on the formulation of the pressure exerted by just one molecule on the given wall; thus it is surely valid for solely one molecule of ideal gas, in
which case N=1.
P
2
2. THE COMPATIBILITY OF THE LAW OF GASES WITH QUANTUM
MECHANICS, BASED ON THE CONSTANCY OF PV  FOR AN ADIABATIC
TRANSFORMATION
There is a relationship satisfying the criteria we have just set; this is the one describing an
adiabatic transformation of gases in a wide temperature range, i.e.
(7)
PV   Constant ,
obtained in the familiar way based on the law of gases, considered together with the first law
of thermodynamics [3], with the usual definition
C
(8)
 P,
CV
where
3
CV  R ,
(9)
2
5
CP  R ,
(10)
2
CV being the heat to be delivered to one mole of ideal gas at constant volume to increase its temperature as much as 1º K, and C P being the heat to be delivered to one mole of
ideal gas at constant pressure to increase its temperature still as much as 1º K. Eqs. (9) and
(10) are exact, for an ideal gas, for which The Boyle Mariotte law [Eq.(1)] holds for even one
molecule. That is if R were replaced by the Boltzmann Constant in these equations, then one
lands at cV  (3/2)k c p  ( 5 /2)k , defined for just one molecule. This means we can confidently use the ratio Cp/Cv or the same, the ratio cp/cv in the quantum world. Furthermore, recall that, Eqs. (9) and (10) are known to be valid, when internal energy levels of molecules are
not excited. Such an assertion is fulfilled for an ideal gas, by definition. And we will find out
that an ideal gas, thus defined, on the basis of the Boyle Mariotte law [Eq.(1)], is in fact, a gas
which is made of non-interacting molecules, each behaving as a simple quantum mechanical
particle locked up (potential energy – wise) in an infinitely high box.
Note that, classically, although it is established that PV  should remain as a constant
quantity, through an adiabatic transformation, no one knew, what this constant would be. No
one would even seemingly wondered whether such a constancy could be explained based on
any universal constant, such as the Planck Constant.
From Eqs. (8) - (10), one has,
5
γ .
(11)
3
We should emphasize that this ratio can well be applied to a gas made of just one molecule, thus quantum mechanically, to a particle in a box.
In an ideal gas, all molecules are considered independently from each other. As we
will see, this can be checked out. Thus, Eq.(7) remains valid, even if the gas consists in just
one molecule. And once again, within the frame of the Kinetic Theory of Gases, one first, is to
express the pressure for one molecule only, before he proceeds for all molecules, making up
the gas, in order to formulate the macroscopic pressure, the gas exerts on the walls of the container (cf. the aforementioned footnote).
It should be stressed that Eq.(7), as expected, embodies, neither the temperature T, nor
the average translational energy E , so we are well off the incorrectly set, dead end problem
we reported above, regarding the search for a pointless bare link between k and h.
Further on, we would like to introduce the following fundamental question:
3
Is Eq.(7) compatible with a corresponding quantum mechanical frame, one would set?
Let us thus consider a particle of mass m with a fixed internal energy state, located in a
macroscopic cube of side L. Herein we will consider the non-relativistic case. Our approach
however can be extended to the relativistic case with no difficulty. The non-relativistic Schrödinger equation furnishes the energy En of the particle in the box, at a given energy level, i.e.


2
2
2
2
2
2
2
ny
n z  h n x  n y  n z
h 2  n x
,
(12)
En 
 2  2 
8m  L2
L
L 
8mL2
where we denoted nx  1, 2, 3, ... , n y  1, 2, 3, ..., nz  1, 2, 3, ... , the quantum numbers to be associated with the corresponding wave function dependencies, on the respective directions x, y
and z. For brevity, we introduced, the subscript “n” which denotes the specific state characterized by the set of integer numbers n x , n y , and n z . [These should of course, not be confused
with the number of moles n (not written in the italic form), introduced in Eq.(1).]
For an ideal gas the “potential energy” within the box, is null. Thus, we have
1
E n  mvn2 ,
(13)
2
vn being the velocity of the particle at the nth energy level.
At the given energy level, the pressure p n exerted by just one particle on either wall,
becomes [cf. Eqs. (4) and (6)]
mv 2 2 En
.
(14)
pn  3n 
3 L3
3L
Now, let us calculate (for just one particle), the product pnV  :

h 2 nx  n y  nz
2
2
2



h 2 nx  n y  nz
2
3 53
8
mL
.
(15)
pnV 
L

3
12m
L3
Hence, this quantity indeed, turns out to be a constant for a given particle of mass m at
the given energy level.
Recall that the total energy En of Eq.(12), ultimately determines the quantized velocity
vn of Eq.(13).
When it is question of many particles instead of just one, normally, we would have
particles at different, possible, quantized states. This is most likely, what leads to the Maxwellian Distribution of particles, with different translational energies, in a container, at a given
temperature. It is on the other hand, this temperature which specifies the average particle. We
can well visualized the average particle, as a single particle, obeying to Eq.(15), thus situated
at the nth level, and of course associate the given temperature with the energy coming into
play, along with Eq.(5).
Not to complicate things, let us get focussed on the average particle, and simply suppose that all others, behave the same. Furthermore, all three components of the average velocity in equilibrium are expected to be the same. Thus, we can rewrite Eq.(15) for the macroscopic pressure Pn exerted at the given average state n, by one mole of gas, on the walls of the
container:†
2

†
 
Rigorously speaking, one must write
2
2
2
NA
h 2 nix  niy  niz
h2 n2
γ
PnV  
 NA
12m
4m
i 1


,
2
2
2
(i)
along with the definition,
4
h2n2
;
(16)
4m
N A is the Avagadro Number.
Eq.(16) well discloses the constant involved by Eq.(7). Note that, at the average state
n (i.e. at the given temperature), the mean square speed of the gas molecules is vn2. The average energy is furnished accordingly, via the framework of Eq.(13). Let us calculate what
would n be, for 1mole of H2, delineating the pressure of 105 Pascal (i.e. 1 atmosphere), in a
volume of 1 m3. From Eq.(16), we obtain:
PnV   N A
10 5 x 12 x 2 x 10 3 x 1839 x 0.9  10 30
n
 23 x 10 8 .
23
2
68
3 x 6.023 x 10 x 6.62 x 10
(17)
3. DISCUSSION
One may question whether or not our approach can be considered to be a general one.
Herein we have simply provided an answer to the enigmatic constant of Eq.(7). If this equation for any reason fails, it is true that, our approach cannot be applied. But if Eq.(7) fails, this
would, before everything else, mean that, the gas at hand is not an ideal gas. Thus we would
expect other effects such as interactions between the molecules making up the gas, coming
into play.
We should like to note that, Eq.(7) is not a relativistic equation, anyway. That is, if ever the constituents of the gas, moved at speeds which can not be neglected as compared to the
speed of light, then, it is not anyway, a valid relationship, and (at the average state denominated by n), it should be replaced, as insinuated by Eqs.(15) and (16), by
(18)
(mγ c 2 )PnV γ  Constant ,
where mγ c 2 then, is the relativistic energy of the average particle, i.e.
c2h2n2
m c  m c 
;
(19)
4L2
Note further that the relationship PnV γ  Constant well holds for a photon gas,
though with the exponent γ  4/3 , instead of 5/3.
We spare the discussion of these interesting points for a future work.
2
2 4
4. CONCLUSION
In this article, we aimed to bridge, classical thermodynamics and quantum mechanics.
Though, we have determined that, toward that aim, it is in vain, to look for a relationship between Bolzmann Constant k, and Planck Constant h. Indeed, a relationship involving both k
and h, such as Eq.(6), is nothing more than a definition of say, the temperature, in terms of
the translational energy of the particle in hand.
1
3N A
 n
NA
i 1
2
ix

2
2
 niy  niz  n 2 .
(ii)
Thus it becomes clear that, if all particles bared the same set of quantum numbers, each with equal quantum
numbers along all three directions, i.e. n x = n y = n z =n, then
n  n2
.
(iii)
5
So, we had to nail down a relationship which involves, neither k, not h, to be able to
work out, our goal. Thus, we came out with the task of working out the constancy of PV  .
The value of the constancy of PV  , is something totally missed over almost a century,
in the literature. As far as records are concerned, no one seems to have even wondered about
the possible value of this constant.
Herein we have calculated this constant, based on quantum mechanics, at last making
a bridge between classical thermodynamics, and quantum mechanics.
Our result further makes that, the behavior of an ideal gas, otherwise defined by the
Boyle Mariotte law [Eq.(1)], is nothing, but a macroscopic manifestation of quantum mechanics, in fact the quantum mechanical behavior of particles, which can all be considered separately, in the same macroscopic box.
Thus, the classical adiabatic constancy of [pressure] x [volume] (5/3) (thus generally, the
frame drawn by the law of ideal gases), happens to be rooted to quantum mechanics, and
seems to be deep. It is that the quantity [mass ] x [pressure] x [volume] (5/3) turns out to be a
Lorentz scalar.
Thereby, we expect this scalar, to be somehow nailed to a Lorentz invariant universal
constant; this constant, more specifically, turns out to be h 2 (the square of the Planck Constant).
Accordingly, for a given mass m, the quantity [pressure] x [volume] (5/3) relates to
h 2 / m ; this is, what we have revealed, in this article.
Henceforth i) the constancy of [pressure] x [volume] (5/3) appears to be an extension of
quantum mechanics to macroscopic scales, but even more essentially, ii) it delineates how the
internal dynamics displayed by a quantum mechanical particle of a given mass, is organized
in conjunction with the size of space, and the dynamics in question takes place in, and this
universally, at all scales [4, 5]. Here, we will not go in any further details of this fundamental
problem.
REFERENCES
[1] L. de Broglie, Annales de Physique (Section II, Chapter 7), 10e Série, Tome III, 1925.
[2] D. Halliday, R. Resnick, J. Walker, Fundamentals of Physics (Chapter 20), John Wiley
& Sons, Inc., 1997.
[3] A. Sommerfeld, Thermodynamics and Statistical Mechanics (Academic Press, NY,
1964).
[4] T. Yarman, Optics and Spectroscopy, Volume 97 (5), 2004 (683).
[5] T. Yarman, Optics and Spectroscopy, Volume 97 (5), 2004 (691).
97 (5), 2004 (691).
[6] D. ter Haar & H. Wergeland, Elements of Thermodynamics, Addison-Wesley, 1960.
[7] P. Souchay, Chimie Générale, Thermodyamique Chimique, Masson et Cie, 1964.
[8] W. J. Moore, H. Aberdam, Physical Chemistry, Prentice Hall, Inc., 1955, New York.
[9] G. J. Van Wylen, R. E. Sonntag, Fundametals of Classical Thermodynamics, John
Wiley & Sons, Inc., 1985.
[10] J. Kestin, J. R. Dorfman, A Course in Statistical Thermodynamics, Academic Press,
New York & London, 1971.
6