Vous êtes sur la page 1sur 39

ssss ttt

sssssssss ttsss
CHAPTER 4

ssssss
P.G. DEPARTMENT OF PHYSICS UNIVERSITY OF KASHMIR
DISSERTATION SUBMITTED IN THE PARTIAL FULFILLMENT FOR THE DEGREE OF M.SC IN PHYSICS
JAHENGIR AHMAD
SHAH
SHAH AARIF UL
ISLAM

TASADUQ HUSSAIN
MIR

SYED PEERZADA
ROUOOF AHMAD
SHAH
Bose-Einstein
Condensation in
External Traps
Under the Supervision of Dr. Sekh Golam Ali

by


2014
U N I V E R S I T Y O F K A S H M I R , H A Z R A T B A L S R I N A G A R - 1 9 0 0 0 6




P0ST GRADUATE DEPARTMENT OF PHYSICS
UNIVERSITY OF KASHMIR,
HAZRATBAL-SRINAGAR-190006, J & K.
____________________________________


CERTIFICATE

This is to certify that the project entitled Bose Einstein Condensation
in External Traps
,,
submitted by Jahengir Ahmad Shah(12069100003),
Shah Aarif Ul Islam (12069100007) , Tasaduq Hussain Mir
(12069100009) and Syed Peerzada Rouoof Ahmad Shah (12069100027)
in the partial fulfillment for the degree of M.Sc in Physics by University
of Kashmir, Hazratbal, Srinagar has been carried out by them under my
consistent guidance and supervision.




Dr. Manzoor A. Malik Dr. Sekh Golam Ali
(Head of The Department ) ( Supervisor )

1 | P a g e


Chapter 1

Introduction

Inclusion of quantum nature of particles in Statistical Mechanics is one of the most important
developments of Modern Physics. It is first started by S N Bose in 1924 in order to derive
Planck's radiation formula which can explain black-body spectrum. In particular, he considers
the radiation as a bunch of photons and finds their distribution in different energy cells with
the assumptions, (i) photons are indistinguishable and (ii) a cell can contain zero to any
number of particles. This distribution of mass-less particles (photons) is generalized by A
Einstein in 1925. The generalized distribution formula, which includes massive as well as
mass-less particles, is known as Bose-Einstein distribution and, the particles which follow
this distribution are called bosons. The statistics behind this distribution is termed as B-E
statistics. The fundamental property of BE statistics, which tells us that a state can
accommodate any number of particles, leads to the concept of Bose-Einstein condensation
when only available state is the lowest energy state.
If a dilute gas of bosons is cooled to temperatures very close to absolute zero (that is, very
near 0 K or273.15 C) then a large fraction of the bosons occupy the lowest quantum state
at which point quantum effects become apparent on a macroscopic scale. This macroscopic
quantum phenomenon is the so-called Bose-Einstein condensation.
Bose-Einstein condensation is based on the indistinguishibility and wave nature of particles,
which are both basic concepts of quantum mechanics. In a simplified picture, atoms in a gas
may be regarded as quantum mechanical wave packets which have an extent on the order of a
thermal de-Broglie wavelength, z
dB
. In particular, z
dB
1 T , where T is the absolute
temperature. At high temperature, the weakly interacting gas can be treated as a system of
Billiard balls. The lower the temperature, the longer is the de-Broglie wavelength. When
atoms are cooled to the point where the thermal de-Broglie wavelength is comparable to the
interatomic separation, then the atomic wave packets overlap and indistinguishibility of the
2 | P a g e

particles becomes important. Fig.1.1.At this moment bosons undergo a phase transition and
form a Bose-Einstein Condensate, the dense and coherent cloud of atoms all occupying the
same quantum mechanical state. The relation between the transition temperature and the peak
density n, can be simply expressed as
nz
dB
3
= 2.612,
where the thermal de-Broglie wavelength is defined as
z
dB
= ( 2n
2
mk
B
I )
1/2

And m is the mass of the atom.















Fig.1.1. Formation of Bose-Einstein condensation
As we mentioned above, that Bose-Einstein condensation, is predicted in 1924. However, it
is realized experimentally in 1995[1,2]. For this great achievement a Nobel Prize in Physics is
also awarded in 2001. Clearly, it takes 70 years to realize BEC experimentally. One of the
main reasons behind this delay is the unavailability of cooling techniques to reach near
3 | P a g e

absolute zero temperature. This difficulty is overcome with the invention of laser cooling and
evaporative cooling techniques.
Our objective in this dissertation is to give a review on the basic concept of Bose-Einstein
condensation starting from Black-body spectrum. In the second Chapter, we discuss black-
body spectrum and Bose's derivation of Planck's formula. In Chapter 3, we discuss Einstein's
generalization of Bose statistics and his prediction of B-E condensation for ideal gas where
there is no trapping. Here, we also calculate transition temperature and critical number of
atoms for BEC to occur. In chapter 4, we consider BEC in a magnetic trap and derive
explicitly the expression for transition temperature and condensate fraction. In addition we
present some cooling and trapping techniques to realize BEC experimentally. We also briefly
outline some properties of trapped BECs and its possible applications. Finally, we give
concluding remarks in chapter 5.















Black-body spectrum and Bose statistics for mass
particles

Explanation of black-body spectrum is one of the
On the basis of classical physics Rayleigh (1900) and Jeans (1905) derive a formula, called
Rayleigh-Jeans formula, that can explain the spectrum in long wavelength region. However,
in the short wavelength region it fails. The formula which can explain whole region of
spectrum is derived by Max Planck in 1900. Here, he considers quantum nature of radiation
in his calculation but does not pay attention to the statistics radiation quanta (photons). In
1924, S N Bose derives Planck's radiation formula considering statistical distribution of
photons and, as a consequence the Bose statistics is invented.
2.1 Black body radiation spectrum
Black-body radiation is the type of
thermodynamic equilibrium with its environment, or emitted by a
non-reflective body) held at constant, uniform temperature. The radiation has a specific
spectrum and intensity that depends only on the tempera

Fig.2.1: Black
body spectrum and Bose statistics for mass
body spectrum is one of the major challenges in late nineteenth century.
On the basis of classical physics Rayleigh (1900) and Jeans (1905) derive a formula, called
Jeans formula, that can explain the spectrum in long wavelength region. However,
ion it fails. The formula which can explain whole region of
spectrum is derived by Max Planck in 1900. Here, he considers quantum nature of radiation
in his calculation but does not pay attention to the statistics radiation quanta (photons). In
Bose derives Planck's radiation formula considering statistical distribution of
photons and, as a consequence the Bose statistics is invented.
2.1 Black body radiation spectrum
is the type of electromagnetic radiation within or surrounding a body in
thermodynamic equilibrium with its environment, or emitted by a black body
reflective body) held at constant, uniform temperature. The radiation has a specific
spectrum and intensity that depends only on the temperature of the body[3]
: Black-body radiation spectrum
4 | P a g e
Chapter 2
body spectrum and Bose statistics for mass-less
major challenges in late nineteenth century.
On the basis of classical physics Rayleigh (1900) and Jeans (1905) derive a formula, called
Jeans formula, that can explain the spectrum in long wavelength region. However,
ion it fails. The formula which can explain whole region of
spectrum is derived by Max Planck in 1900. Here, he considers quantum nature of radiation
in his calculation but does not pay attention to the statistics radiation quanta (photons). In
Bose derives Planck's radiation formula considering statistical distribution of
surrounding a body in
black body (an opaque and
reflective body) held at constant, uniform temperature. The radiation has a specific
[3] (Fig 2.1).

5 | P a g e


Black-body radiation has a characteristic, continuous frequency spectrum that depends only
on the body's temperature. The spectrum is peaked at a characteristic frequency that shifts to
higher frequencies with increasing temperature, and at room temperature most of the
emission is in the infrared region of the electromagnetic spectrum. As the temperature
increases past about 500 degrees Celsius, black bodies start to emit significant amounts of
visible light. Viewed in the dark, the first faint glow appears as a "ghostly" grey. With rising
temperature, the glow becomes visible even when there is some background surrounding
light: first as a dull red, then yellow, and eventually a "dazzling bluish-white" as the
temperature rises[4]. When the body appears white, it is emitting a substantial fraction of its
energy as ultraviolet radiation. The Sun, with an effective temperature of approximately 5800
K, is an approximately black body with an emission spectrum peaked in the central, yellow-
green part of the visible spectrum, but with significant power in the ultraviolet as well.

2.2 Rayleigh-Jean formula
To explain the black-body spectrum, Rayleigh and Jeans begins the classical calculation by
considering a blackbody as a radiation filled cavity at the temperature T (Fig 2.2)

Fig.2.2: A hole in the wall of a hollow object is an excellent approximation of a blackbody.


6 | P a g e


Fig.2.3: EM radiation in a cavity whose walls are perfect reflectors consists of standing waves that have nodes
at the walls, which restricts their possible wavelengths. Shown above are three possible wavelengths when the
distance between the opposite walls is L.
Since the cavity walls are assumed to be perfect reflectors, the radiation must consist of
standing electromagnetic waves, as in (Fig 2.3). In order for a node to occur at each wall, the
path length from wall to wall, in any direction, must be an integral number j, of half-
wavelengths. If the cavity is a cube, having edge of length L, this condition means that for
standing waves in the x, y, and z directions respectively, the possible wavelengths are such
that

]
x
=
2L
x
, ]

=
2L
x
, ]
z
=
2L
x
. (2.1)
Here ]
x
, ]

and ]
z
represent the number of half-wavelengths in the x,y and z directions
respectively and they can take any integer value, e.g. 1,2,3 and so on.
For a standing wave in any arbitrary direction, it must be true that in order the wave terminate
in a node at its ends
]
x
2
+]

2
+]
z
2
= [
2L
x

2
. (2.2)
Here, ]
x
= u,1,2, ]

= u,1,2, and ]
z
= u,1,2
7 | P a g e

(Of course, if ]
x
= ]

= ]
z
= u, there is no wave, though it is possible for any one or two of
the ]s to equal 0.)
To count the number of standing waves g(z)Jz within the cavity whose wavelengths lie
between z and z +Jz, what we have to do, is to count the number of permissible sets
]
x
, ]

, ]
z
values that yield wavelengths in this interval. Lets imagine a j-space whose
coordinate axis are ]
x
, ]

and ]
z
. Fig.2.4 shows part of the ]
x
]

plane of such a space. Each


point in the j-space corresponds to a permissible set of ]
x
, ]

, ]
z
values and thus correspond to
a standing wave.

Fig.2.4: Each point in the j-space corresponds to a possible standing wave.
If j is a vector from the origin to a particular point ]
x
, ]

, ]
z
its magnitude is
] = ]
x
2
+]

2
+]
z
2
(2.3)
The total number of wavelengths between z and z +Jz is the same as the number of points
in j-space whose distance from the origin lie between ] and ] +J]. The volume of the
spherical shell of radius ] and thickness J] is 4n]
2
J], but we are only interested in the octant
of this shell that includes non negative values of ]
x
, ]

and ]
z
. Also, for each standing wave
counted in this way, there are two perpendicular directions of polarization, hence the no, of
independent standing waves in the cavity is
g(])J] = (2) [
1
8
(4n]
2
J]) = n]
2
J]. (2.4)
What we really want is the no. of standing waves in the cavity as a function of their
frequency v instead of as a function of ]. From equations (2.2) and (2.3) we have

8 | P a g e

] =
2L
x
=
2Lv
c
onJ J] =
2L
c
Jv. (2.5)
Therefore, no. of standing waves
g(v)Jv = n [
2Lv
c

2
2L
c
Jv =
8nL
3
c
3
v
2
Jv. (2.6)
The cavity volume is I
3
, which means that the number of independent standing waves per
unit volume or the density of standing waves in a cavity is
0(v)Jv =
1
L
3
g(v)Jv =
8nv
2
dv
c
3
. (2.7)
Equation (2.7) is independent of the shape of the cavity, even though we used a cubical cavity
to facilitate the derivation. The higher the frequency, the shorter the wavelength and greater
the number of standing waves that are possible.
The next step is to find the average energy per standing wave. According to the classical
theorem of equipartition of energy, as already mentioned, the average energy per degree of
freedom of an entity that is part of a system of such entities in thermal equilibrium at the
temperature T is
1
2
kI. Each standing wave is a radiation-filled cavity corresponds to two
degrees of freedom, for a total e of kI, because each wave originates in an oscillator in the
cavity wall. The energy per unit volume, u(v)Jv in the cavity, in the frequency interval from
v to v +Jv is therefore, according to classical physics,
u(v)Jv = e0(v)Jv = kI0(v)Jv =
8nv
2
k1dv
c
3
, (2.8)
which is Rayleigh-Jeans formula[5].
The Rayleigh-Jeans formula, which has the spectral energy density of blackbody radiation
increasing as v
2
without limit, is obviously wrong. Not only does it predicts a spectrum
different from the observed one (see Fig.2.5), but integrating (2.8) from v = u to v = give
the total density as infinite at all temperatures.


9 | P a g e


Fig.2.5: Comparison of Rayleigh-Jeans Formula for the spectrum of the radiation from a blackbody at 1500K
with the experimentally observed spectrum. The discrepancy is known as Ultraviolet catastrophe because it
increases with increasing frequency.
This discrepancy between theory and observation was at once recognized as fundamental.
This is the failure of classical physics that led Max Plank in 1900 to discover that only if light
emission is considered as a quantum phenomenon, can correct the formula for u(v)Jv be
obtained. And thus we arrive at the Planks radiation formula.

2.3 Planks Radiation Formula
Plank found that he had to assume that the oscillators in the cavity walls were limited to
energies of e
n
= nv, where n = u,1,2, . He then used the Maxwell-Boltzman distribution
law to find that the number of oscillators with the energy e
n
is proportional to c
-c
n
k1
at the
temperature T. In this case the average energy per oscillator (and so per standing wave in the
cavity) is
e =
hv
c
hv kT
-1
(2.9)
Instead of the energy-equipartition average kI which Rayleigh and Jeans had used. . Here is
where Classical and Quantum physics diverge. The energy in the interval v to v +Jv now
becomes
10 | P a g e


u(v)Jv = e0(v)Jv =
8nh
c
3
v
3
dv
c
hv kT
-1
(2.10)
which is Planks Radiation formula[5], that agrees with the experimental findings.
To check, how does Planks formula explain experimental results, we consider different
cases, as:
(i) When v kI, the term c
hv k1
1 v kI ,
hence equation (2.10) can be written as
u(v)Jv =
8nk1
c
3
v
2
J v. (2.11)
Which is Rayleigh Jeans law and explains the blackbody radiation at longer wavelengths and
higher temperatures.
(ii) Whenv kI, the eq. (2.10) becomes
u(v)Jv =
8nhv
3
c
3
c
-hv k1
Jv. (2.12)
Which is Weins law and explains the blackbody radiation spectrum at shorter wavelengths
and lower temperatures. Thus Planks radiation law successfully explains the blackbody
spectrum at shorter as well as at longer wavelengths.

2.4. Boses derivation of Planks Radiation Formula
Planks formula for the distribution of energy in a blackbody radiation forms the starting
point for the quantum theory and has yielded rich harvests in all fields of physics. Since its
publication in the year 1901, many types of derivations of this law have been suggested. It is
acknowledged that the fundamental assumptions of quantum theory are inconsistent with the
laws of classical electrodynamics. All existing derivations make use of the following relation
p
v
Jv = (8nv
2
Jv c
3
)E, (2.13)
representing the relation between radiation density and the mean energy of an oscillator.
11 | P a g e

Let the radiation be enclosed in a volume V and its total energy be E. Let their be different
species of quanta each characterized by the no. N
s
and energy v
s
(s = u s = ).The
total energy E is then
E = N
s
v
s s
= I p
v
Jv (2.14)
The solution of our problem requires then the determination of the numbers N
s
which
determine p
v
. If we can state the probability for any distribution characterized by an arbitrary
set of N
s
,then the solution is determined by the requirement that the probability be a
maximum provided auxiliary condition (2.14) is satisfied. It is this probability which we now
intend to find.
The quantum has a moment of magnitude v
s
c in the direction of its forward motion. The
instantaneous state of the quantum is characterised by its coordinates x, y, z and associated
momenta p
x
, p

, p
z
. These six quantities can be interpreted as point coordinates in a six
dimensional space; they satisfy the relation
p
x
2
+p

2
+p
z
2
=
2
v
2
c
2
(2.15)
By virtue of which the above mentioned point is forced to remain on a cylindrical surface
which is determined by the frequency of the quantum. In this sense the frequency domain Jv
s

is associated with the phase space domain
JxJyJzJp
x
Jp

Jp
z
= I4n(v c )
2
Jv c = 4n(
3
v
2
c
3
)IJv (2.16)
If we subdivide the total phase-space volume into cells of magnitude
3
, then the no.of cells
belonging to the frequency domain Jv is
4nI(v
2
c
3
)Jv (2.17)
Concerning the kind of subdivision of this type nothing definitive can be said. However, the
total no. Of cells must be interpreted as the no. of the possible arrangements of one quantum
in the given volume. In order to take into account the polarization, it appears mandatory to
multiply this no. by the factor 2 so that the no. of cells belonging to an interval Jv becomes
8nI(v
2
c
3
)Jv (2.18)
It is now very simple to calculate the thermodynamic probability of a macroscopically
defined state. Let N
s

be the no. of quanta belonging to the frequency domain Jv
s
. In how
12 | P a g e

many different ways can we distribute these quanta over those cells which belong to the
frequency interval Jv
s
.
Let p
0
s
be the no. of vacant cells, p
1
s
the no. of those cells which contain one quantum, p
2
s
the
no. of cells containing two quanta, etc; then the no. of different distributions is

A
s
!
p
0
s
! p
1
s
!
(2.19)
where A
s
= (8nv
2
c
3
)Jv
s

and N
s
= up
0
s
+1p
1
s
+2p
2
s
+
is the no. of quanta belonging to the interval Jv
s
.
The probability of the state which is defined by all the p

s
is obviously

A
s
!
p
0
s
!p
1
s
!
s

In view of the fact that we can look at the p

s
as large numbers, we have
ln w = A
s
ln A
s
p

s
lnp

s
s s
(2.20)
where

A
s
= p


This expression should be maximum satisfying the auxiliary condition
E = N
s
v
s
;
s
N
s
= rp

(2.21)
Carrying out the variation gives the condition
op

s
(1 +ln p

s
) = u, oN
s
v
s
= u,
s s
(2.22)
op

s
= u, oN
s
= rop

s

(2.23)
It follows that
op

s
(1 +ln p

s
+z
s
) +
1
[
v
s
s
rop

= u
s
(2.24)
From (2.24), we can write
13 | P a g e

p

s
= B
s
cxp [
hv
s
[
. (2.25)
Using (2.25) in

A
s
= p

,
we get.
A
s
= B
s
cxp [
hv
s
[
= B
s
j1 cxp [
hv
s
[
[
-1

(2.26)
we have
B
s
= A
s
j1 cxp [
hv
s
[
[ (2.27)
Furthermore we have the relation
N
s
= rp

= rA
s
j1 cxp [
hv
s
[
[

cxp [
hv
s
[
=
A
s
cxp(-hv
s
[ )
1-cxp(-hv
s
[ )
(2.28)
Because of the above stated value of A
s
, it is also true that
E =
8nhv
s
3
dv
s
c
3
s
I
cxp(-hv
s
[ )
1-cxp(-hv
s
[ )
(2.29)
Using the preceding results, one finds also that
S = k ]
L
[
A
s
lnj1 cxp [
hv
s
[
[
s
. (2.30)
From this it follows that [ = kI, because of the condition

oS
oE
, =
1
I
, . (2.31)
Substituting kT for [ in above equation for E, one obtains
E =
8nhv
s
3
c
3
I jcxp [
hv
s
k1
1[
-1
Jv
s
s
, (2.32)
which is equivalent to the Planks formula[6]. The function within the square brackets is
equivalent to energy per mode in Planks derivation.


14 | P a g e

Chapter 3

Bose-Einstein Condensation for an ideal gas

Bose's derivation of Plank's formula includes statistics of photons which are massless. It fails
to include statics of massive particles having same properties of photons. Just after Bose's
work, Einstein gives a general statistics to include massive particles also. This statistics is
known as Bose-Einstein statistics and the particles behaving in accordance to Bose's statistics
are today called bosons. Bosons in three-dimensional configuration start to condense in the
lowest energy state when they are cooled below a critical temperature .

3.1 Bose-Einstein Statistics
In order to discuss Einsteins generalization of Bose statistics, we consider a system of N
bosons with total energy U. Suppose that the system has an energy level e

with
degeneracy g

, containing n

bosons. The states may be represented by g

1 lines, and the


bosons by n

circles; distinguishable microstates correspond to different orderings of the lines


and circles. For example, with 9 particles in 8 states corresponding to a particular energy, a
particular microstate might be considered as shown in Fig.3.1:

Fig.3.1: Distribution of particles in different states
The number of distinct orderings of lines and circles using 2.19 is
t

=
(n

+g

1)!
n

! (g

1)!
(S.1)
A particular distribution has a specified number of particles n

within each of the possible


energy levels e

. The total number of microstates for a given distribution is therefore:


15 | P a g e

t({n

]) = _
(n

+g

1)!
n

! (g

1)!

. (S.2)
Let us assume that each state has a high degeneracy, i.e.g

1. Then we can make the


approximation:
t({n

]) _
(n

+g

)!
n

! g

. (S.S)
To find the most probable distribution, we maximise (3.3):
t({n

]) = _
(n

+g

)!
n

! g


subject to the constraint on the total number of particles:
n

= N (S.4)
and the constraint on the total energy:
e

= u (S.S)
As usual, rather than maximise t directly, we maximise ln t. If we assume that both g


and n

are large enough for Stirlings approximation to hold for lng

! and ln n

! we find that
ln t is given by:
ln t |(n

+g

) ln(n

+g

) g

ln g

Jn

lnn

. (S.6)
The change in ln t resulting from changes Jn

in each of the populations n

is then:
J ln t |(n

+g

)Jn

ln n

Jn

(S.7)
From the constraints (3.4) and (3.5), we find:
Jn

= u

, e

Jn

= u (S.8)
16 | P a g e

Combining (3.7) and (3.8) with Lagrange multipliers and , we have:
J lnt _ln _
n

+g

] +o +[e

Jn

. (S.9)
For appropriate values of and , equation (3.9) is true for all Jn

, hence we have
ln _
n

+g

] +o +[e

= u. (S.1u)
We then find that the most probable distribution can be written as
n

=
g

c
-u-[s
i
1
. (S.11)
Equation (3.11) is the Bose-Einstein distribution[7]. It gives the population of an energy level
that has energy e

and degeneracy g

. The constants and are determined from the


constraints (3.4) and (3.5) on the total number of particles and the total energy. can, as
usual, be related to the thermodynamic temperature, so that the Bose-Einstein distribution
with c
-u
= B takes the form:
n

=
g

Bc
s
i
k1
1
. (S.12)
This distribution implies that we can have n

, if e

kI, and B 1. In other words, it


is possible for many identical bosons to exist in the same state. This curious behaviour
displayed by bosons leads to the phenomenon of Bose-Einstein condensation.
It is important to note that in the beginning it was unknown about the nature of the particles
which will obey this statistics. After a few years it became clear that the Boson should have
integer spin and the wave function for a system of identical bosons must be symmetric with
respect to the interchange of any two of the bosons. For example, with two bosons, a possible
wave function might be:
(x
1
, x
2
) =
1
2
|
A
(x
1
)
B
(x
2
) +
A
(x
2
)
B
(x
1
)] (S.1S)
If we put two particles into the same state, the wave-function does not vanish: there is no
limit on the number of particles we can put into any given state. Therefore, properties of
particles which follow the Bose-Einstein distribution is justified.
17 | P a g e


3.2 Bose-Einstein Condensation
Equation (3.12) shows that n

is directly proportional to g

. As a result
n

=
1
Bc
s
i
k1
1
(S.14)
may be interpreted as the most probable number of particles per energy level in the i th cell. It
is important to note that the final result (3.14) is totally independent of the manner in which
the energy levels of the particles are grouped into cells so long as the number of levels in
each cell is sufficiently large. Also, the entropy of the gas is given by
S
k
= |n

(o +[e

) g

ln|1 c
-u-[s
i
|]

= oN +[E g

ln|1 c
-u-[s
i
|


(S.1S)
From(3.15)
S
k
= [ _
o
[
N +E
1
[
g

ln|1 c
-u-[s
i
|

_ (S.16)
Equation (3.16) in conjunction with the second law of thermodynamics gives [ =
1
k1

and o =

k1
, where is the chemical potential defined as p = [
L
N

v,1
.
As we noted, the result in (3.12) is totally independent of the manner in which the energy
levels of the particles are grouped into cells so long as the number of levels in each cell is
sufficiently large. In view of this, using the value of and in (3.12) we get the mean
occupation number of the level e in the form.
n
s
=
1
c
(s-) k1
1
. (S.17)
where e denotes the energy of the single-particle state.
From (3.17) it is clear that, for the mean occupation number to be positive, < e. When
becomes equal to the lowest value of e, say e
mn,
the occupancy of that particular level
18 | P a g e

becomes infinitely high. Since the number of particles is conserved, the chemical potential p
enters in the distribution function (3.17).
The chemical potential is determined as a function of N and T by the condition that the total
number of particles be equal to the sum of the occupancies of the individual levels. It is
sometimes convenient to work in terms of the quantity = c

k1
,
, which is known as the
fugacity. If we take the zero of energy to be that of the lowest single-particle state, the
fugacity is less than unity above the transition temperature and equal to unity (to within terms
of order 1/N, which we shall generally neglect) in the condensed state. In Fig. (3.2) the
distribution function (3.17) is shown as a function of energy for various values of the
fugacity.
At high temperatures, the effects of quantum statistics become negligible, and the
distribution function (3.17) is given approximately by the Boltzmann distribution. At high
temperatures the chemical potential lies well below e
mn
, the energy of the lowest single-
particle state, since in this limit the mean occupation number of any state is much less than
unity, and therefore, in particular, cxp|(p e
mn
)kI] 1. As the temperature is lowered,
the chemical potential rises and the mean occupation numbers increase. However, the
chemical potential cannot exceed e
mn
, otherwise the Bose distribution function (3.17)
evaluated for the lowest single-particle state would be negative, and hence unphysical.
Consequently the mean occupation number of any excited single-particle state cannot exceed
the value 1 {cxp j
(c
v
-c
min
)
k1
[ 1] , . If the total number of particles in excited states is less than
N, the remaining particles must be accommodated in the single-particle ground state, whose
occupation number can be arbitrarily large: the system has a BoseEinstein condensate[8].
The highest temperature at which the condensate exists is referred to as the BoseEinstein
transition temperature and we shall denote it by
c
. The energy dependence of the single-
particle density of states at low energies determines whether or not BoseEinstein
condensation will occur for a particular system. In the condensed state, at temperatures below

c
, the chemical potential remains equal to e
mn
, to within terms of order kIN, which is
small for large N, and the occupancy of the single-particle ground state is macroscopic in the
sense that for N , a non-zero fraction of the particles are in this state. The number of
particles N
0
in the single-particle ground state equals the total number of particles N minus
the number of particles N
ex
occupying higher-energy (excited) states.
19 | P a g e


Fig. 3.2: The Bose distribution function
0
= 1(
-1
cxp(ekI) 1) as a function of energy for different
values of the fugacity . The value = 1 corresponds to temperatures below the transition temperature, while =
0.5 and = 0.25 correspond to 0.69kT and 1.39kT, respectively.

3.3 Critical Temperature and Number of Atoms
We know that all particles in the gaseous states can go into the lowest energy state at
extremely low temperature. At higher temperatures gaseous particles obey classical statistics
of Maxwell and Boltzmann given by
n
s

M.B
= c
(-s) k1
. (S.18)
From (3.17), we can go (3.18) provided
c
(s-) k1
1. (S.19)
Equation (3.19) tells us that the chemical potential of the system must be negative. Thus the
fugacity, z = c
kT
of the system must be smaller than unity. The quantity z reflects the
tendency of a substance to prefer one phase to another and can literally be defined as the
tendency to escape.
Moreover, in a quantum mechanical theory z is related to N and V by
N
I
= z
(2nmkI)
3 2

3
. (S.2u)
20 | P a g e

In terms of thermal de Broglie wave length
z =

2nmkI
, (S.21)
Combining (3.20) and (3.21), we can write
N
I
=
z
z
3
(S.22)
From (3.22), for z to be less than unity, we must have
z
3
N
I
1. (S.2S)
The quantity nz
3
(n = N/V ) or fugacity is an appropriate parameter in terms of which the
various physical properties of the system can be addressed. For example, we consider three
cases. (i) nz
3
0. In that case, 0 and the particle aspect of the gas molecules or atoms
dominates over the wave aspect. Thus the system is classical. (ii) 1 > nz
3
> 0. We can now
expand all physical quantities as a power series in this parameter and investigate how the
system tends to exhibit non-classical or quantum behaviour. (iii) nz
3
1. The system
becomes significantly different from the classical one and typical quantum effects begin to
dominate.
From (3.21)
nz
3
=
nh
3
(2nmk1)
3 2
(S.24)
This expression clearly shows that the system is more likely to display quantum behaviour
when it is at a relatively low temperature or has a relatively high density of particles.
Moreover, for smaller particle mass, the quantum effects will be more prominent.
From (3.17) the total number of particles N in the system is obtained as
N = n
s

s
=
1
z
-1
c
[s
1
s
(S.2S)
For large volume V , the spectrum of the single-particle state is almost a continuous one,
summation on the right hand side of (3.25) may be replaced by integration. The density of
states (e) in the neighbourhood of e is given by
21 | P a g e

p(e)Je =
2nI

3
(2m)
3 2
e
1 2
Je (S.26)
so that
N
I
=
2n

3
(2m)
3 2
_
e
1 2
z
-1
c
[s
1
+

0
1
I
z
1 z
. (S.27)
In writing (3.27) we have separated out the e = 0 term in (3.25) which has a statistical weight
equal to one. Denoting z/(1 z ) by N
0
we write (3.27) in the form
N N
0
I
=
2n

3
(2nmkI)
3 2
_
x
1 2
z
-1
c
x
1

0
(S.28)
with x = e. In terms of Bose-Einstein functions,
b
v
(z) =
1
(u)
_
x
v-1
Jx
z
-1
c
x
1

0
= z +
z
2
2
v
+
z
3
S
v
+ (S.29)
the result in (3.28) can be written as
N N
0
I
=
1
z
3
g
3 2
(z). (S.Su)
The quantity( N N
0
) denotes the number of particles (N
e
) in the excited states. Therefore,
N
c
= I _
2nmkI

2
]
3 2
b
3 2
(z). (S.S1)
The function b
3 2
(z) increases monotonically and is bounded with largest value
b
3 2
(1) = 1 +
1
2
3 2
+
1
S
3 2
+ = (S 2 ) = 2.612 (S.S2)
Hence, for all z, of interest
b
3 2
(z) (S 2 ). (S.SS)
In view of (3.33), N
e
in (3.31) will satisfy the condition
N
c
I _
2nmkI

2
]
3 2
(S 2 ). (S.S4)
22 | P a g e

The equality in (3.34), gives the maximum number of particles in the excited states. If the
actual number of particles N of the system exceeds this limiting values, then N
0
number of
particles given by
N
0
= N I _
2nmkI

2
]
3 2
(S 2 ) (S.SS)
will be pushed into the ground state. Since N
0
= z/(1 z ), the precise value of z can be
determined using
z =
N
0
N
0
1
1. (S.S6)
For z to be one, the chemical potential must be zero. Thus from (3.17),
n
s
=
1
exp(e kI ) 1
. (S.S7)
This result shows that for large N, there is no limitation to the number of particles that can go
into the ground states e = 0. This curious phenomenon of a macroscopically large number of
particles accumulating in a single particle states e = 0 is referred to as Bose-Einstein
condensation. It is purely of quantum mechanical origin, even in the absence of inter-particle
forces. It takes place in the momentum space.
The condition for the onset of Bose-Einstein condensation is
N > N
e
(3.38)
which gives a critical value of temperature as[9]
I
c
=

2
2nmk
_
N
I(S 2 )
]
2 3
. (S.S9)
For given values of N and V, Bose-Einstein condensation takes place when temperature T of
the gas is less than I
c
.



23 | P a g e

Chapter 4

Bose-Einstein Condensation in External Traps

In the 1980s laser based techniques were developed to trap and cool neutral atoms (Chu,
Cohen Tannoudji, Phillips). Technically, trapping and cooling in this approach go by the
names, magneto-optical trapping (MOT) and laser cooling. Alkali metal atoms are well suited
to laser based methods because their optical transitions can be excited by available lasers and
because they have a favourable internal energy-level structure for cooling to very low
temperatures. Once they are trapped, their temperature can be lowered further by evaporative
cooling to observe Bose-Einstein Condensation.
4.1 Effect of trapping
With a view to calculate effects of trapping, we first calculate density of state for both the
free and trapped gas. While calculating density of states for a trapped BEC, we shall assume
that all particles are in one particular internal (spin) state, and therefore we generally suppress
the part of the wave function referring to the internal state. In most cases the confining traps
are well approximated by harmonic potentials, and therefore, we consider only particles in
harmonic trap.
In three dimensions, for a free particle in a particular internal state, there is on average one
quantum state per volume (2n)
3
of phase space. The region of momentum space for which
the magnitude of the momentum is less than p has a volume 4np
3
S equal to that of a sphere
of radius p and, since the energy of a particle of momentum p is e
p
= p
2
2m, the total
number of states G(e) with energy less than e is given by[8]
G(e) = I
4n
3
(2mc)
32
(2n)
3
= I
2
13
3n
2
(mc)
32

3
(4.1)
where V is the volume of the system. Quite generally, the number of states with energy
between e and e+d e is given by g(e)d e, where g(e) is the density of states. Therefore
g(e) =
du(c )
dc
(4.2)
24 | P a g e

which, from Eq. (4.1), is thus
g(e) =
vm
32
2
12
n
2

3
e
12
. (4.3)
For a free particle in d dimensions, the density of states is independent of energy for a free
particle in two dimensions.
Let us now consider a particle in the anisotropic harmonic-oscillator potential (trap)
V (r) =
1
2
m(K
x
x
2
+K

y +K
z
z) (4.4)
which we will refer to as a harmonic trap. Here the quantities K

(i = x, y, z) denote the three


force constants, which are generally unequal. The corresponding classical oscillation
frequencies
i
are given by

2
=
K

m

and we shall therefore write the potential as
V (r) =
1
2
m(
x
2
x
2
+

2
y +
z
2
z) (4.5)
The energy levels, e(n
x
, n

, n
z
), are then
e(n
x
, n

, n
z
)=(n
x
+
1
2
)
x
+[n

+
1
2

+[n
z
+
1
2

z
(4.6)
where the numbers n

assume all integer values greater than or equal to zero.


We now determine the number of states 0(e) with energy less than a given value e . For
energies large compared with

, we may treat the n

as continuous variables and neglect


the zero-point motion. We therefore introduce a coordinate system defined by the three
variables e

, interms of which a surface of constant energy (4.6) is the plane


e = e
x
+e

+e
z
. Then 0(e) is proportional to the volume in the first octant bounded by the
plane,
0(e) =
1

3
o
x
o
j
o
z
Je
x
Je

Je
z
=
c
3
6
3
o
x
o
j
o
z
c-c
x
-c
j
0
c-c
x
0
c
0
(4.7)
Sinceg(e) = J0Je, we obtain a density of states given by
25 | P a g e

g(e) =
c
2
2
3
o
x
o
j
o
z
(4.8)
For a d-dimensional harmonic-oscillator potential with frequencies

, the analogous result is


g(e) =
c
d-1
(d-1)! o
i i
. (4.9)
We thus see that in many contexts the density of states varies as a power of the energy, and
we shall now calculate thermodynamic properties for systems with a density of states of the
form
g(e) = C
u
e
u-1
(4.10)
whereC
u
is a constant. In three dimensions, for a gas confined by rigid walls, is equal to
3/2. The corresponding coefficient may be read off from Eq. (4.3), and it is
C
32
=
vm
32
2
12
n
2

3
(4.11)
Understandably, (4.3) for d=3, coincides with (4.10) for o =3/2.
The coefficient for a three-dimensional harmonic-oscillator potential ( = 3), which may be
obtained from Eq. (4.8), is
C
3
=
1
2
3
o
x
o
j
o
z
(4.12)
4.1.1 Critical temperature
The transition temperature I
c
is defined as the highest temperature at which the macroscopic
occupation of the lowest-energy state appears. When the number of particles, N, is
sufficiently large, we may neglect the zero-point energy in (4.6) and thus equate the lowest
energy e
mn
to zero, the minimum of the potential (4.5).
The number of particles in excited states is given by
N
cx
= J(e)g(e)
0
(e)

0
. (4.13)
Where
0
(e) = n(e), as given in eq. (3.37)
This achieves its greatest value for = 0, and the transition temperature I
c
is determined by
the condition that the total number of particles can be accommodated in excited states, that is
26 | P a g e

N = N
cx
(I
c
, p = u) = Jcg(e)
1
c
ekT
c
-1

0
(4.14)
When (4.14) is written in terms of the dimensionless variable x = ekI
c
, it becomes
N = C
u
(kI
c
)
u
Jx
x
o-1
c
x
-1

0
= C
u
(o)(o)(kI
c
)
u
(4.15)
Where (o) is the gamma function and (o) = n
-u
n=1
is the Riemannzeta function. In
evaluating the integral in (4.15), we expand the Bose function in powers of c
-x
, and use the
fact that Jxx
u-1

0
c
-x
= (o). The result is

x
o-1
c
x
-1
=

0
(o)(o) (4.16)
Table 4.1 lists (o) and (o) for selected values of o.

Table 4.1 The gamma-function and the Rieman-zeta function for selected values of o
From (4.15) we now find
kI
c
=
N
1o
|C
o
I(u)(u)]
1o
(4.17)
For a three-dimensional harmonic-oscillator potential, o is 3 and C
3
is given by Eq. (4.12).
From (4.17) we then obtain a transition temperature given by
kI
c
=
oN
13
|(3)]
13
u.94N
13
(4.18)
where
= (
x

z
)
13
(4.19)
27 | P a g e

is the geometric mean of the three oscillator frequencies. The result (4.18) may be written in
the useful form
I
c
4.S[
]

100Hz
N
1
3
nK, (4.20)
where

= 2n.
For a uniform Bose gas trapped in a three-dimensional box of volume V the index o is 3/2.
Using the expression (4.11) for the coefficient C
32
, one finds for the transition temperature
the relation
kI
c
=
2n
|(32)]
23

2
n
23
m
S.S1

2
n
23
m
(4.21)
I
c
=
2n
|(S 2 )]
2 3

2
n
2 3
m
. (4.22)
where n = NI is the number density. For a uniform gas in two dimensions, o is equal to 1,
and the integral in (4.15) diverges. Thus BoseEinstein condensation in a two-dimensional
box can occur only at zero temperature. However, a two-dimensional Bose gas can condense
at non-zero temperature if the particles are confined by a harmonic-oscillator potential. In
that case o = 2 and the integral in (4.15) is finite.
When the number of particles is extremely high we can neglect the zero point energy in the
harmonic trap. This is, however, not true when the system consists of finite number of atoms.
The finiteness of the number of particles calls for zero point energy to be taken into account.
This reduces the critical temperature by an amount T
c

such that

I
c
I
c
=
(2)
2|(S)]
2 3

N
-1 3
, (4.2S)
where
m
=
1
3
(
1

3
)
1 3
.Clearly from (4.23)
A1
c
1
c
u, as N . Thus we see that one
of the effects of trapping is to lower the critical temperature by confining a finite number of
atoms. Besides finiteness of the system, trapping makes the Bose gas inhomogeneous such
that density variations occur on a characteristic length scale, o
ho
= (m
m
) , provided by
the frequency of the trapping oscillator. This is a major difference with respect to other
systems like the super fluid helium where the effects of in-homogeneity take place on a
microscopic scale in the coordinate space. In-homogeneity of super fluid helium, in fact
cannot be detected in the coordinate space such that all observations are made in the
28 | P a g e

momentum space. As opposed to this, the in-homogeneity of the Bose gas is such that both
coordinate and momentum spaces are equally suitable for observations.

4.1.2 Condensate Fraction
Below the transition temperature the number N
cx
of particles in excited states is given by
N
cx
(I) = C
u
Jee
u-1

0
1
c
ekT
-1
, (4.24)
provided the integral converges, that is o > 1. We may write this result as
N
cx
= C
u
(o)(o)(kI)
u
, (4.25)
where C
u
is a constant. In three dimensions, for a gas confined by rigid walls, is equal to
3/2, so that
C
32
=
Im
32
2
12
n
2

3
(4.26)
Note that this result does not depend on the total number of particles. However, if one makes
use of the expression for I
c
, it may be rewritten in the form
N
cx
= N_
I
I
c
]
u
(4.27)
The number of particles in the condensate is thus given by
N
0
(I) = N N
cx
(I) (4.28)
or
N
0
= Nj1 [
1
1
c

u
[ (4.29)
For particles in a box in three dimensions, o is 3/2, and the number of excited particles per
unit volume, obtained from Eqs. (4.26) and (4.27) is given by
29 | P a g e

n
cx
=
N
cx
v
= (S2) [
mk1
2n
2

32
(4.Su)
The condensate fraction is therefore given by
N
0
= N|1 (II
c
)
32
] (4.31)
For a three-dimensional harmonic-oscillator potential (o = S), the condensate fraction is
given by

N
0
N
= _1 _
I
I
c
]
3
_ (4.S2)
Comparing 4.31 and 4.32 one can see the variation of condensate fraction in the two cases.

4.2 Method of Cooling and Trapping
Basically laser and evaporative cooling techniques are used to produce BEC in the
Laboratory. In order to trap the gas, one can use magnetic as well as electric fields. Here we
consider only magnetic trap.
4.2.1 Laser cooling
Laser beams are often used to pre-cool the atomic vapour and the method used goes by the
name laser cooling. The physical mechanism by which the collision between photons and
atoms reduces the temperature of the atomic vapour can be visualized as follows.
If an atom travels toward the laser beam and absorbs a photon from the laser it will be
slowed down by the photon impact. Understandably, totality of such events will lower the
temperature. On the other hand, if the atom moves away from the photon, the latter will speed
up resulting in the increase of temperature. Thus it is necessary to have more absorption from
head on photons if our goal is to slow down the atoms with a view to lower the temperature.
One simple way to accomplish this in practice is to tune the laser slightly below the
resonance absorption of the atom.
30 | P a g e

Suppose that the laser beam is propagating in a definite direction. An atom in the gaseous
system can move towards the beam or it may move away from the beam. In both cases the
frequency of the photon will be Doppler shifted. In the first case the frequency of the laser
beam will increase while in the other case the frequency will be decreased. In the case of
head on collision the photon will be absorbed by the atom via resonance only when the
original laser beam is kept below the frequency of atomic resonance absorption. When the
atom and photon travel in the opposite direction, there cannot be momentum transfer from the
photon to the atom because Doppler shift in this case produces further detuning of the already
detuned laser beam. The minimum temperature achieved so far by laser cooling technique is
less than 10
-6
K.

Fig.4. 1: Laser cooling

4.2.2 Magnetic Trapping
Magnetic traps are used to confine low-temperature atoms produced by laser cooling. These
traps use the same principle as that in the Stern-Gerlach experiment. Otto Stern and Walter
Gerlach used the force produced by a strong inhomogeneous magnetic field to separate the
spin states in a thermal atomic beam as it passes through the magnetic field. But for cold
atoms the force produced by a system of magnetic coils bends the trajectories right around so
that low energy atoms remain within a small region close to centre of the trap. This can be
realized as follows.
A magnetic dipole momentp in a magnetic field B

has energy
I = p. B

. (4.SS)
For an atom in a hyperfine state |I[FH
P
B, V corresponds to a Zeeman energy
I = g
P
p
B
H
P
B, (4.S4)
where p
B
= Bohr magneton and
31 | P a g e

g
P
g
]
F(F +1) +[([ +1) S(S +1)
2F(F +1)
(4.SS)
The magnetic force along z-direction is
P = g
P
p
B
H
P
JB
Jz
(4.S6)
From (4.33) the energy depth of the magnetic trap is determined by p

B. The atomic
magnetic moment p

is of the order of Bohr magneton p


B
which in temperature units 0.67
Kelvin/Tesla. Since laboratory magnetic fields are generally considerably less than 1 Tesla,
the depth of magnetic traps is much less than a Kelvin, and therefore atoms must be cooled in
order to be trapped magnetically.
For confinement, Zeeman energy must have a minimum. We can consider two different cases
for (4.34).
Case 1 : H
P
g
P
> u. Here the Zeeman energy can be minimum, if B has a local minimum.
Case 2 : H
P
g
P
< u. In this case V can have a local minimum if B has a local maximum.
Maxwells equations do not allow a maximum of a static field. As a result the trapping of
atoms for H
P
g
P
< u is not allowed. In view of the above one can trap atoms only in a
minimum of a static magnetic field. This explains how magnetic field can be used to trap
atoms. For detail we refer the book in [10].

4.2.3 Evaporative cooling and condensates density profile
The basic idea of the evaporative cooling is simple. First, let's consider atoms trapped by the
magnetic potential as shown in the panel (a) of the picture below. The most energetic (i.e.
fastest, hottest) atoms, in the course of their movement within the trap, can go much higher
up the potential walls. Now, we lower the walls of the trap potential. As a result, the most
energetic atoms will likely fall out of the trap. The remaining atoms will collide with each
other, exchanging momentum and reaching a new, lower temperature equilibrium; this is
called re-thermalization. After the atoms have been allowed to re-thermalize, we lower the
walls of the trap potential again, and allow the atoms to rethermalize again. This procedure is
repeated until the critical temperature of a BEC is reached[11,12].
32 | P a g e


Fig. 4.2: Evaporative cooling
4.3 Experimental Observation of BEC
An observation Bose-Einstein condensation in the laboratory for different values of
temperatures, namely, below, at and above critical temperatures are shown in Fig. 4.3.









Fig.4.3: Velocity distribution of Rubidium-87 condensates. The false colors indicate the number of atoms at
each velocity, with red being the fewest and white being the most. The areas appearing white and light blue are
at the lowest velocities.
The Bose-Einstein condensate is characterized by its slow expansion observed 6 ms after the
atom trap was turned off. The left picture shows an expanding cloud cooled just above the
critical temperature; middle: just after the condensate appeared; right: after further
evaporative cooling has left an almost a pure condensate.
In the first observations [13], four features were used to identify the formation of a Bose-
Einstein condensate:

(1) The sudden increase in the density of the cloud.
33 | P a g e


(2) The sudden appearance of a bimodal cloud consisting of a diffuse normal component

(3) The velocity distribution of the condensate was anisotropic in contrast to the isotropic

expansion of the normal (non-condensed) component.

(4) The good agreement between the predicted and measured transition temperatures
Possible applications of Bose-Einstein Condensate:
Atomic BEC systems are in basic research areas. It is supposed that the BEC will have a
lot of applications in near future. (i) Bose condensates can be used to simulate condensed
matter systems. For example, In the presence of "optical lattice", a periodic potential created
by interference pattern of multiple laser beams[14], the atoms behave like electrons in
crystal lattices. The big advantage of BEC optical lattice systems over real condensed matter
systems is that lattice parameters are more easily tunable. (ii) A possible application of BEC
is its use in precision measurement. (iii) Researchers are looking for ways to use BEC
systems for quantum information processing. (iv) It is also hoped to use of BECs in atomic
clocks and atom lasers [15].















34 | P a g e


Chapter 5

Conclusion
Studies of Bose-Einstein condensation in external potentials have received a great deal of attention. Since
the experimental realization of BEC in 1995, a huge amount of work has been done in this direction[16].
In view of this, we review in this dissertation some interesting and important topics related to Bose-
Einstein condensation. We start from Black-body spectrum and discuss how does Bose's concept of
photon distribution gives the Planck's radiation formula. We also discuss how Einstein's generalization
leads to BE statistics. One of the important consequences of BE statistics is occurrence of BE
condensation at an extremely low temperature. We consider Bose gases in the presence and absence of
external potentials separately and give a detail review. Here we clearly mentioned effects of trapping on
the critical temperature and number of atoms. In addition, we show by simple calculations that an external
trapping allows condensation to occur even in low dimensional systems. We also briefly discuss different
cooling and trapping techniques that are required to realize BEC in the laboratory. Characteristics of
ultra-cold dilute Bose gas after the onset of BECs are also pointed out. We conclude by noting that the
dissertation serves as a good introductory review in a very straightforward manner.











35 | P a g e



Bibliography:

[1] K.B. Davis, M.-O. Mewes, M.R. Andrews, N.J. van Druten, D.S. Durfee, D.M. Kurn, W.
Ketterle,BoseEinstein condensation in a gas of sodium atoms, Physical Review
Letters 75,3969 (1995)
[2] M.R. Anderson, J.R. Ensher, M.R. Matthews, C.E. Wieman, E.A. Cornell, Observation of
Bose-Einstein Condensation in a Dilute Atomic Vapour, Science 269, 198 (1995)
[3] P. Theodore Landsberg, Thermodynamics and statistical mechanics, (Reprint of Oxford
University Press 1978 ed.) Courier Dover Publications. (1990).
[4] J.W. Draper, On the production of light by heat, London, Edinburgh and Dublin
Philosophical Magazine and Journal of Science, 30: 345 (1847).
[5] A. Beiser, S. Mahajan and S.R. Choudhury, Concepts of Modern Physics , McGraw Hill
Education (India) ,2013.
[6] S.N. Bose, Planks Law and Light Quantum hypothesis, Zeitschrift fr Physik ,
26,178(1924)
[7] F. Reif, Fundamentals of Statistical and thermal Physics, McGraw Hill Book Company
[8] C. J. Pethick and H. Smith, BoseEinstein Condensation in Dilute Gases, Cambridge
University Press, Cambridge, 2001.
[9] R.K. Pathria and P. D.Beale, Statistical Mechanics, Elsevier, 2011.
[10] H. J. Metcalf, P.van der Straten, Laser Cooling and Trapping, Springer New York,1999.
[11] K.B. Davis, M.-O. Mewes, W. Ketterle, An analytical model for evaporative coolingof
atoms, Appl. Physics. B 60, 155 (1995).
[12] K.B. Davis, M.-O. Mewes, M.A. Joffe, M.R. Andrews, W. Ketterle, Evaporative
coolingof sodium atoms, Physical Review Letters 74,5202 (1995).
36 | P a g e

[13] W. Ketterle, M.R. Andrews, K.B. Davis, D.S. Durfee, D.M. Kurn, M.-O. Mewes,
N.J.van Druten, Bose-Einstein condensation of ultracold atomic gases, Phys. Scr. T66,
31(1996).
[14] I. Boch, Ultracold Quantum gases in Optical lattices, Nature Physics 1,23 (2005).
[15] M.H. Anderson, J.R. Ensher, M.R. Mathews, C.E. Wieman, E.A. Cornel, Observation of
Bose-Einstein Condensation in a dilute Atomic Vapour, Science 269, 198(1995).
[16] S. G. Ali, M. Salerno, A. Saha and B.Talukdar, Displaced dynamics of binary mixtures
in linear and non-linear optical lattices, Physical Review A 85,023639(2012).

Vous aimerez peut-être aussi