You are on page 1of 25

QUARKONIUM STATES

WITH POTENTIAL
MODELS

Project Report by

Khitish Biswal
1411048
School of Physical Sciences
NISER

under the guidance of

Dr. Subhasis Basak


CONTENTS
INTRODUCTION 1

1. THE SCHRÖDINGER EQUATION 2

2. FEYNMAN HELLMAN THEOREM 3

2.1. APPLICATIONS OF FEYNMAN HELLMAN THEOREM 3

3. VIRIAL THEOREM 4

4. MASS DEPENDENCE OF SCHRÖDINGER WAVEFUNCTIONS 8

5. POWER LAW AND LOGARITHMIC POTENTIAL 11

5.1. DEPENDENCE UPON MASS AND COUPLING STRENGTH 12

5.2. THE INFINITE SQUARE WELL 15

5.3. LOGARITHMIC POTENTIAL 16

6. SEMI-CLASSICAL APPROXIMATIONS 18

7. APPLICATIONS TO QUARKONIUM 18

7.1. MASS DEPENDENCE OF OBSERVABLES 18

7.2. QUANTUM NUMBER DEPENDENCE OF OBSERVABLES 19

7.3. LOGARITHMIC POTENTIAL 20

7.4. CORNELL POTENTIAL 22

8. CONCLUSION 23
INTRODUCTION
In this article we look at the quark anti-quark bound states , also known as quarkonium, of
charm and bottom quarks. We generally solve Dirac’s equation to describe light particles
which includes quarks, since major part of the energy in it is in the form of momentum it
calls for a relativistic treatment. The lighter quarks are in the order of few MeVs so it can’t
be treated non relativistically. However with increasing quark mass like for the charmed
quark which has a mass of 1.29 GeV, the non-relativistic approximation becomes better. The
Dirac’s equation will anyway be a more accurate description, but its not very easy to solve
it. Thus, we revert to Schrödinger’s equation which is relatively easier to solve. To solve the
Schrödinger’s equation we need a potential to work with. We start by looking at some
potentials which are basically the power law potentials. We also look at the logarithmic
potential whose behaviour can be approximated to a power law. Then we take some physical
behaviour of the potential known to us under account. We know that the quark has a charge,
thus the potential would contain a coulomb like term. Also, we have that the quark and
antiquark are confined. So the potential would contain a term proportional to some power of
r, where r is the distance between the two. We look at the simplest potential where the
second term in the potential is linear in r . Thus, we get a potential of the form,
V(r) = − a /r + br, which is also know as the Cornell potential, a and b are the parameters
that are determined by fitting the results of the calculations to the masses of well-measured
quarkonium states.
This report starts with the general consequences of the Schrödinger’s equation. Then
we acquire some tools, that will be essential in the later sections, which includes the
Feynman-Hellmann theorem and the generalized virial theorem. In the subsequent section
we use the tools to get results which holds for potentials of the power law form and the
logarithmic potential. After this, we take some results obtained from semiclassical
approximations and then we turn our attention to the quarkonium spectroscopy. There we
study two specific potentials for the charmonium and the upsilon families and try to predict
the mass of a higher energy state with the aid of the results obtained and inputs from the
experimental results of the lower energy states.


1
1. THE SCHRÖDINGER EQUATION

The Schrödinger equation in three dimensions is

ℏ2 2
− ∇ ψ + [V(r) − E ]ψ = 0 (1.1)

where μ is the reduced mass of the two-body system and r is the relative coordinate, ψ is the
wavefunction and V is the interaction potential. For a central potential, we can write

ψ = R(r)Ylm(θ, ϕ) (1.2)

where R(r) is the radial wavefunction and Ylm is a spherical harmonics which is the angular
wavefunction.

Substituting ψ in \\ the radial part of the Schrödinger equation reads

2μ ( dr 2 r dr ) [ 2μr 2 ]
ℏ2 d2 2 d l(l + 1)ℏ2 (1.3)
− + R(r) − E − V(r) − R(r) = 0

This equation can be reduced to one dimensional problem with the substitution u(r) ≡ rR(r)

ℏ2 [ 2μr 2 ]
2μ l(l + 1)ℏ2
−u′′(r) = E − V(r) − u(r) (1.4)
satisfying the boundary conditions

u(0) = 0 (1.5a)
u′(0) = R(0) (1.5b)

The Schrödinger wavefunction is normalised


2
dV ψ =1
(1.6)


and dΩ Y* Y = δll′δmm′
lm l′m′
(1.7)

The reduced radial wavefunction satisfies


dr [u(r)] = 1

2 (1.7)

2
2. FEYNMAN HELLMAN THEOREM

The Feynman Hellman theorem describes the relationship of the derivative of the total
energy E with respect to some parameter λ with the expectation value of the derivative of
the Hamiltonian (H ) with respect to the same parameter λ.

Let ψ (λ) be an eigenvector of H(λ) with eigenvalue E(λ). We can write the time independent
Schrödinger equation as

H(λ) | ψ (λ)⟩ = E(λ) | ψ (λ)⟩ (2.1)

let ψ (λ) be normalised i.e. ⟨ψ (λ) | ψ (λ)⟩ = 1 (2.2)


or ⟨ψ (λ) | ψ (λ)⟩ = 0 (2.3)
∂λ

The theorem states that

∂λ ⟨ ∂λ ⟩
∂E ∂H
= (2.4)

The theorem is proved as following,

E(λ) = ⟨ψ (λ) | H(λ) | ψ (λ)⟩ (2.5)


∂E ∂
∂λ
= ∂λ
⟨ψ (λ) | H(ψ) | ψ (λ)⟩ (2.6)
∂⟨ψ (λ) ∂H(ψ) ∂ψ (λ)
= ∂λ
| H(ψ) | ψ (λ)⟩ + ⟨ψ (λ) | ∂λ
| ψ (λ)⟩ + ⟨ψ (λ) | H(λ) | ∂λ

= ⟨ ∂∂λH ⟩ + E
∂⟨ψ |
∂λ
ψ⟩ + E⟨ψ | ∂λ
∂λ

= ⟨ ∂∂λH ⟩ + E ∂
∂λ
(⟨ψ | ψ⟩)

= ⟨ ∂∂λH ⟩ + E ⋅ 0

= ⟨ ∂∂λH ⟩ (Q.E.D)

2.1. APPLICATIONS OF FEYNMAN HELLMAN THEOREM

Given Hamiltonian H(λ) with eigenvalues En(λ) and normalised eigenstates ψn(λ)

The generalised force can be defined as

Fλ ≡ − ∂λ En(λ)
(2.7)
= − ∂λ⟨H ⟩
3
For our interest we look at the variation of bound state energy with respect to the reduced
mass.

Given the Hamiltonian of the form

H(μ) = T (μ) + V(r)


−ℏ2 ∇2
(2.8)
=− 2μ
+ V(r)

Taking the variation with respect to μ, we get


∂H 1
∂μ
= (ℏ2∇2 )
2μ 2
1 ℏ2 2
= ( ∇)
2μ μ 2
(2.9)
= − μ1 (H(μ) − V(r))

= ⟨ ∂H
∂μ ⟩
∂E
∂μ
(2.10)
= − μ1 (E − ⟨V ⟩)

From this, we can see that the energy of a particular bound state decreases with increase in
μ.

3. VIRIAL THEOREM
The virial theorem relates the mean value of the kinetic energy to the gradient of the
potential. We shall derive the virial theorem from a more general expression.

We shall write the Schrödinger equation in the following way for ease

−u′′(r) = ℓ(r)u(r) (3.1)

where ℓ(r) is defined as

ℏ2 [ 2μr 2 ]
2μ l(l + 1)ℏ2 (3.2)
ℓ(r) = E − V(r) −

Multiplying r qu′(r) on both sides and integrating from 0 to ∞ , we get


∞ ∞

∫0 ∫0
dr r qu′(r)u′′(r) = dr qℓ(r)u(r)u′(r) (3.3)

4
1
u′u′′ = ((u′)2 )′ (3.4)
2
1 2 (3.5)
uu′ = (u )′
2

Substituting back in the equation


∞ ∞

∫0 ∫0
dr r q(u′(r)2 )′ = dr r qℓ(r)[u(r)2]′ (3.6)

Doing integration by parts for the left hand side becomes


∞ ∞

∫0 ∫0
dr r q(u′(r)2 )′ = − r q[u′(r)]2 |∞
0 +q dr r q−1[u′(r)]2 (3.7)

the second term of the equation becomes


∞ ∞ ∞
q ∫0 r q−1[u′(r)]2 dr = qr q−1u(r)u′(r) − q ∫0 dr r q−1u(r)u′′(r) − q(q − 1) ∫0 dr r q−2u(r)u′(r)

= qr q−1u(r)u′(r) |0 + q⟨r q−1ℓ(r)⟩ − q(q − 1) ∫0 dr r q−2u(r)u′(r) (3.8)

Here the first term vanishes at infinity since both u and u′ vanishes at infinity for a bound
state

The remaining integral is again evaluated by parts to get



q q(q − 1)(q − 2) q−3
∫0
−q(q − 1) dr r q−2u(r)u′(r) = − (q − 1)r q−2[u(r)]2 |0 + ⟨r ⟩ (3.9)
2 2

Now, evaluating the right hand side of the equation


∞ ∞
∫0 dr r qℓ(r)[u(r)2]′ = r qℓ(r)[u(r)]2 |0 + q ∫0 dr r q−1[u′(r)]2 (3.10)
∞ ∞
= r qℓ(r)[u(r)]2 |0 − q ∫0 dr r q−1ℓ(r)[u(r)]2 − ∫0 dr r qℓ′(r)[u(r)]2

Here we again used the fact that for the bound states the value for the upper limit of
integration vanishes since u and u′ vanishes at infinity.

Rewriting the original equation after evaluation of the integrals we get,

[r q[u′(r)]2 − qr q−1u(r)u′(r) + 12 q(q − 1)r q−2[u(r)]2 − r qu(r)(u′′(r)] |r=0


1 (3.11)
= − 2q⟨r q−1ℓ(r)⟩ − ⟨r qℓ′(r)⟩ − q(q − 1)(q − 2)⟨r q−3⟩
2

5
Here we take a digression and look at behaviour of u(r) near the origin.

From the radial part of the Schrödinger equation in three dimensions, we have

ℏ2
− u′′ + Vef f u = Eu (3.12)

l(l + 1)ℏ2
where , Vef f = V(r) + (3.13)
2μr 2

We claim that u(r) → 0 as r → 0

We will prove it by contradiction. Assume lim u(r) = c, where c is any constant. Now the
r→0
full wave function ψ is the product of both wavefunctions.

u(r)
ψ (r) = Y , where Ylm is the wave function of the angular part. (3.14)
r lm

Ylm is a constant for l = 0 . So,


c (3.15)
ψ ≃
r

Putting it back in the Schrödinger equation we get

ℏ2 2
∇ ψ + Vef f ψ = E ψ (3.16)

We know that,

(r)
1 (3.17)
∇2 = 4πδ 3(r)

c
So the Schrödinger equation is not satisfied for ψ ≃ near the origin since the Vef f term
r
does not have a δ function to cancel this term.

So,
lim u(r) = 0 (3.18)
r→0

Near the origin we will look at the solutions for u(r) with the conditions u(0) = 0 and
u′(0) = R(0) (3.19)

6
d 2u 2m
=− Eu + Vef f u (3.20)
dr 2 ℏ2
l(l + 1)
≃ u
r2
(3.21)
u = al r l+1 + bl r −l

For l = 0, the second term is inconsistent with the Schrödinger equation at r = 0.


For l > 0, the second term is not consistent with normalisation.

So, we have u(r) = al lr l+1 for r → 0. (3.22)

u′ = al (l + 1)r l and u′′ = al (l + 1)(l )r l−1 (3.23)

Now substituting u′ and u′′ back in the left side of the equation //

r q(al (l + 1)r l )2 − qr q−1al r l+1al (l + 1)r l + 12 q(q − 1)r q−2


1
= al2(l + 1)2r q+2l − al2 q(l + 1)r q+2l + q(q − 1)al2r q+2l − al2r q+2l l(l + 1) (3.24)
2

In the limit r → 0, setting q = − 2l to get non-trivial solutions to the equation

al2[(l + 1)2 + 2l(l + 1) + 12 (−2l )(−2l − 1) + l(l + 1)] = = al2[l 2 + l + 1 + 2l 2 + 2l + 2l 2 + 2l − l 2 − l ]


= al2[4l 2 + 4l + 1] (3.25)

So, rewriting eqn. 3.11 we arrive at,

1
(2l + 1)2 al2 δq,−2l = − ⟨2qr q−1ℓ + r qℓ′ + q(q − 1)(q − 2)r q−3⟩, which is true for any
2
power q ≥ − 2l. (3.26)

Now, we will look at some results that can be deduced from this equation. Taking l = 0 and
q = 0, we get

a02 = − ⟨−ℓ′(r)⟩

⟨ dr ⟩
2μ dV (3.27)
=
ℏ2

u′(0) = a0 = R(0) = 4π ψ (0) (3.28)

2π ℏ2 ⟨ dr ⟩
μ dV
| ψ (0) |2 = (3.29)

Taking q = 1 and l = 0,
7
0 = − ⟨2ℓ(r) + rℓ′(r)⟩
(3.30)
⟨2ℓ(r)⟩ = − ⟨rℓ′(r)⟩

By rearranging, we get

⟨ 2 ( dr )⟩
1 dV
E − ⟨V ⟩ ≡ ⟨T ⟩ = r , which is the virial theorem. (3.31)

For power law potentials, i.e.


(3.32)
V(r) = λr ν

⟨T ⟩ = 12 ν⟨V ⟩
νE (3.33)
= 2+ν

or

2E (3.34)
⟨V ⟩ =
2+ν

4. MASS DEPENDENCE OF SCHRÖDINGER WAVEFUNCTIONS

We shall deal with one dimensional bound state problem for a symmetric potential, and then
extend it to the three dimensional central potential problem. In this section we will try to
establish the dependence of the observables with mass μ.

We start by examining the probability P(R) that a particle be contained within the interval 0
to R. In terms of the normalised wave function u(x),
R ∞

∫0 ∫0
2
P(R) ≡ [u(x)] d x ,where [u(x)]2 d x = 1 (4.1)
We will establish for symmetric potential with V′(x) > 0

We know that,
0 ≤ P(R) ≤ 1 and P′(R) ≥ 0 i.e. P(R) is an increasing function of R.

Writing down the one dimensional Schrödinger equation,


−2μ
u′′(x) = 2 [E − V(x)]u(x) (4.2)


Now operating u(x) on both sides,
∂μ

8
[ ℏ2 ]
−2μ

(E − V(x))u(x)
∂u′′(x)
u(x) = u(x) (4.3)
∂μ ∂μ

Assuming V(x) to be independent of μ

ℏ [ ∂μ ] [ ] ∂μ
∂u′′(x) −2 ∂E 2μ ∂u(x)
u(x) = 2 E − V(x) + μ [u(x)]2 − 2 [E − V(x)] (4.4)
∂μ ℏ

From the Feynman-Hellman theorem, we know that

∂E 1
= − [E − ⟨V ⟩]
∂μ μ

Using this eqn. 4.4 can be written as


∂u′′(x) −2 2μ ∂u(x)
u(x) ∂μ
= [E − V(x) + μ[ −1
μ
(E − ⟨V ⟩)]][u(x)]2 − [E − V(x)] ∂μ
ℏ2 ℏ2
2 2μ ∂u(x)
= [V(x) − ⟨V ⟩][u(x)]2 + [V(x) − E ]u(x) ∂μ
(4.5)
ℏ2 ℏ2

∂u(x)
Multiplying on both sides of eqn. 4.2 we get
∂μ

∂μ ( ℏ2 )
∂u ∂u −2μ
u′′(x) = [E − V(x)u(x)] (4.6)
∂μ

Subtracting this from the above equation we get,

∂u′′(x) ∂u 2
u(x) − u′′(x) = 2 [V(x) − ⟨V ⟩][u(x)]2 (4.7)
∂μ ∂μ ℏ

Now integrating this from 0 to R


R R
∂u′′(x) ∂u 2
∫0 ∫
(u(x) − u′′(x))d x = [V(x) − ⟨V ⟩][u(x)]2 d x (4.8)
∂μ ∂μ 0 ℏ
2

( ∂μ )]
2 R
[ ∂μ
∂u′ ∂u′ ∂u ∂u′
∫ ∂μ ∫ ∂μ ℏ ∫0
u − u′ dx − u′ − u′d x = 2 d x[V(x) − ⟨V ⟩][u(x)]2 (4.9)
0

To evaluate the left hand side we will discuss about how to apply the boundary conditions.

For a one dimensional case a wave function must be of a defined parity i.e. either
ψ (x) = ψ (−x) or ψ (x) = − ψ (−x) if the potential function is even. This arises from the fact
that for a symmetric potential in one dimension both ψ (x) and −ψ (x) satisfies the time
9
independent Schrödinger equation. So the eigenvector of the hamiltonian must be of defined
parity.

From Born’s conditions we know that the wave function and its derivative must be
continuous.

1. For odd-parity states


(4.10)
ψ (x) = − ψ (−x), or ψ (x) = 0

ψ (x + h) − ψ (x) (4.11)
ψ′(x) = lim
h→0 h

At, x = 0

ψ (h) − ψ (0)
ψ′(0) = lim approaches a finite quantity.
h→0 h

So, we get the boundary conditions for odd parity


ψ (x) = 0 and ψ′(x) is finite at x = 0.

2. For even-parity states


ψ (x) = ψ (−x), ψ can be any finite quantity. Thus, ψ (x) is finite and ψ′(x) = 0 at x = 0.

So the limit at x = 0 vanishes for the left hand side of the eqn. 4.9 and we are left with

∂u′(R) ∂u(R) 2 R
ℏ ∫0
u(R) − u′(R) = 2 [V(x) − ⟨V ⟩][u(x)]2 d x ≤ 0 . (4.12)
∂μ ∂μ
The inequality arises from the fact that V′(x) ≥ 0

Now we define a function G (R),


R
1 ∂P(R) ∂u(x)
∫0
G (R) = = d x u(x) (4.13)
2 ∂μ ∂μ

∂u(R) ∂u
G′(R) = u(R) , which vanishes only if u = 0 or =0 (4.14)
∂μ ∂μ

∂u(R) ∂u′(R) (4.15)


G′′(R) = u′(R) + u(R)
∂μ ∂μ

Evaluating G′′(R) at the extrema of the function we get

∂u(R)
G′′(R) = u′(R) > 0 if u(R) = 0 (a minimum) (4.16)
∂μ

10
∂u′(R) ∂u(R)
G′′(R) = u(R) < 0 if = 0 (a maximum) (4.17)
∂μ ∂μ

Since ,
G (0) = 0 = G (∞),

the function G (R) can become negative for 0 < R < ∞ only if it has a minimum in this
domain. But we proved above that the minimum occurs only at the node of the
wavefunction i.e. u(x) = 0 . Since the ground function wavefunctions are nodeless ,
G (R) ≥ 0 for 0 ≤ R < ∞.

We just established that G (R) ≥ 0 , 0 ≤ R < ∞ for the odd parity case. By applying the
same method to the three dimensional central potential case we can extend the proof to all
nodeless wavefunctions for arbitrary values of the angular momentum l.

Instead of the mass μ we see the variation κ where κ is the potential strength we arrive at a
similar result.

5. POWER LAW AND LOGARITHMIC POTENTIAL


We will now look at some results of the power law potentials i.e. potential of the form
V(r) = λr ν, −2 < ν < ∞ , the domain for ν is based upon the fact that ⟨T ⟩ ≥ 0 , and the

( r0 )
r
logarithmic potential V(r) = C ln , which can be approximated as a power law

potential with ν → 0, λ → ∞ and λν → C.

Fig. 1: Behaviour of logarithmic potential and approximated power law potential

11
From the virial theorem,

⟨ 2 ( dr )⟩
1 dV
⟨T ⟩ = E − ⟨V ⟩ = r

For power law,

⟨T ⟩ = 12 ⟨r (νλr ν−1⟩
(5.1)
= 12 νV

⟨T ⟩ = E − ⟨V ⟩
(5.2)
2⟨T ⟩
=E− ν
or

νE (5.3)
⟨T ⟩ =
2+ν

and similarly for the logarithmic potential


(5.4)
C
⟨T ⟩ =
2

5.1. DEPENDENCE UPON MASS AND COUPLING STRENGTH

We will perform some scaling operations in the reduced radial Schrödinger equation to
make it dimensionless.

To begin with the reduced radial Schrödinger equation

[ 2μr 2 ]
ℏ2 ν l(l + 1)ℏ2
u′′(r) + E − λr − u(r) = 0 (5.5)

We scale r and E in the following way

Let,
r = aρ (5.6)

d d dρ
dr
= dρ dr
1 d (5.7)
= a dρ

(5.8)
12
d 1 d
=
dr 2 a 2 dρ

Rewriting the Schrödinger equation in terms of ρ, we get

[ ℏ2 a ρ ]
1 2μ 2μ ν ν l(l + 1) (5.9)
u′′(ρ) + E − λ a ρ − 2 2 u(r) = 0
a2 ℏ2

[ ( ℏ2 ) ρ2 ]
2μ 2μ 2+ν ν l(l + 1) (5.10)
u′′(ρ) + a 2 E − λa ρ − u(ρ) = 0
ℏ2

To make the equation dimensionless we set

( | λ | 2μ )
2μ 2+ν ℏ2 2+ν
λ a = 1 or a = (5.11)
ℏ2

Scaling βE = ϵ (5.12)

β = a2 (
ℏ2 )
2μ (5.13)

( )
= | λℏ| 2μ 2 + ν ⋅ ( 2 )
2 2μ
(5.14)

−2

( | λ | 2μ ) ( 2μ )
2
ℏ 2+ν ℏ2
E= ⋅ ϵ (5.15)

Defining w(ρ) ≡ u(r)

The Schrödinger equation in the dimensionless form becomes

[ ρ2 ]
l(l + 1)
w′′(ρ) + ϵ − sgn(λ) ⋅ ρ ν − w(ρ) = 0 (5.16)

5.1.1. LEVEL SPACINGS


We can shift the energy eigenvalues by adding a constant to the potential. However the
energy eigenvalue differences are given as

13
−2

( | λ | 2μ ) ( 2μ )
ΔE ∝ ℏ2 2+ν ⋅ ℏ2
(5.17)

−ν 2
∝ ( 2 )2 + ν |λ|2 + ν

(5.18)

Now we look at the dependence of the energy eigenvalue differences on mass for some
specific potentials

Coulomb potential (ν = − 1) :
ΔE ∝ μ | λ |2 , We can verify this using hydrogen atom energy which is linearly dependent on
the mass and dependent on the square of the coupling constant (Z for the hydrogen like
atoms case).
Harmonic oscillator(ν = 2)
1

(μ)
λ 2
ΔE ∝ (5.19)

This can also be verified from the fact that for the harmonic oscillator the energy levels
k
ΔE ∼ ℏω, where ω =
μ

For the linear potential (ν = 1) which is used to describe quark confinement,


1

(μ)
2
λ 3
ΔE ∝ . (5.20)

Also we saw that the behaviour of the logarithmic potential can be approximated as (ν → 0)
which tells us that the energy eigenvalue differences will be independent of mass.

5.1.2. LENGTHS
−1
According to eqn. 5.11 the quantities with dimension of length varies as L ∝ (μ | λ | ) 2 + ν

We can see that the probability densities which have a dimension of L −3 will have the
dependence
3
| ψ (0) |2 ∝ (μ | λ | ) 2 + ν (5.21)

Thus, for the coulomb potential | ψ (0) |2 ∝ (μ | λ | )3. For the linear potential

| ψ (0) |2 ∝ (μ | λ | ). (5.22)
14
The dependence of the probability of the ground state wavefunction for the linear potential
can also be verified from eqn. 3.26

1
(2l + 1)2 al2 δq,−2l = − ⟨2qr q−1ℓ + r qℓ′ + q(q − 1)(q − 2)r q−3⟩
2
for l = 0, q = 0

a02 = − ⟨ℓ′(r)⟩
2μ dV
= ⟨ dr ⟩
ℏ2

Since,

u′(0) = a0 = R(0) = 4π ψ (0)

2π ℏ2 ⟨ dr ⟩ 2π ℏ2
μ dV μλ
| ψ (0) |2 = = (5.23)

5.2. THE INFINITE SQUARE WELL


The infinite square well potential of width R can be thought of as the approximation of
power law with the limit ν → ∞, and the coupling strength is given by R −ν.

ν→∞ ( R )
ν
r
V(r) = lim (5.24)

For this potential, the energy eigenvalue dependence upon mass and coupling strength is
given by
−ν 2

( ℏ2 )
2μ 2 + ν
ΔE ∝ ⋅ |λ|2 + ν (5.25)

−ν 2
lim ΔE ∝ ( 2 ) 2 + ν ⋅ | λ | 2 + ν

ν→∞ ℏ
−1
2 2
+1
∝ ( 2) ν

⋅ (R −ν ) 2 + ν , Subst it ut ing for λ = R −ν

∝(
ℏ2 )
2μ −1
⋅ R −2 (5.26)

15
This result can be verified from the energy eigenvalues of the infinite square well potential

n 2ℏ2 π 2
En = (5.27)
2m L 2

5.3. LOGARITHMIC POTENTIAL

As we saw above that the logarithmic potential can be approximated as the ν → 0, λ → ∞


and λν → C of the power law potential.
−ν 2

( ℏ2 )
2μ 2 + ν
ΔE ∝ ⋅ |λ|2 + ν (5.28)

Putting ν = 0 , we arrive at the result that the energy eigenvalue differences for the
logarithmic potential is independent of the mass.

To prove this result we rescale the Schrödinger equation for ν = 0 and take a general
potential V(r) = λU(r). The Schrödinger equation now becomes

(λ) ( 2μλ )
2
E ℏ 2 l(l + 1)
w′′(ρ) + −U ⋅ρ− w(ρ) (5.29)
ρ2

The eigenvalue differences will be independent of mass iff under a scale transformation
μ → σμ

1 1

[ 2σμλ ] [ 2μλ ]
2 2
ℏ 2 ℏ 2
U ⋅ ρ =U ⋅ ρ + f (σ) (5.30)

1
U(r) = ln(r) satisfies the above equation with f (σ) = − ln(σ) . Thus for the logarithmic
2
potential the level spacings are independent of μ . The parameter λ sets the scale of the
E
spacings, since the dimensionless energy eigenvalue is ϵ = .
λ

16
In this section we looked upon the variation of energy levels upon mass and the coupling
strength for the power law potentials. We approximated the logarithmic and the square well
potential to a power law and verified that the results hold true for them too.

From Feynman Hellman theorem we can see that the levels fall deeper into the potential as
μ is increased.
∂E
∂μ
= − μ1 (E − ⟨V ⟩)

V = − 1/ r V =r

Fig. 2: Energy level spacings

Fig. 3: ΔE vs m

17
6. SOME RESULTS TAKEN FROM SEMI-CLASSICAL
APPROXIMATION ( WKB )
r
For the logarithmic potential V(r) = C ln( ) is singular at r = 0
r0

l 1
2 π (n+ − )
Enl = C ln 2 4
(6.1)
1
2μC
( 2 ) 2 r0

For Power law potentials, the energy eigenvalues are given by


| Enl | ∼ (n − γl (ν)) 2 + ν (6.2)

where,
−l /2, for ν = ∞
1 − 2l
for 0 < ν < ∞ (6.3)
γl (ν) = 4
ν + 1 − 2l
for  − 2 < ν < 0
2(2 + ν)

7. APPLICATIONS TO QUARKONIUM
In this section we will try and deduce the effective power for the power law potentials
through various results derived in the previous sections.

7.1. MASS DEPENDENCE OF OBSERVABLES

7.1.1 LEVEL SPACINGS


As we saw in Section 5.1.1
−ν
ΔE ∝ μ 2 + ν (7.1)

mb = 9.460 GeV
mc = 3.096 GeV, where mb and mc denotes Υ mass and J/ψ mass respectively. (7.2)

mb
= 3.056, (7.3)
mc

18
Using the relation 6.2
(E2s ) Υ − (E1s ) Υ 10.023 − 9.460
(E2s ) ψ − (E1s ) ψ
= 3.686 − 3.096
5.629
= 5.891 (7.4)
= 0.954
ΔEΥ −ν m
log = log Υ (7.5)
ΔEψ 2+ν mψ

−ν log 0.954
2+ν
= log 3.056
(7.6)
= −0.042

ν = 0.080 (7.7)

This gives the effective power for the potential of the power law form.

7.2. QUANTUM NUMBER DEPENDENCE OF OBSERVABLES

7.2.1 LEVEL SPACINGS


From the relation from the semi-classical approximation, we have


| Enl | ∼ (n − γl (ν)) 2 + ν

for l = 0 state

0, for ν = ∞
1
for 0 < ν < ∞
γl (ν) = 4
ν +1
for  − 2 < ν < 0
2(2 + ν)

From experimental results we have

ψ family Υ family
E30 − E20 0.58 ± 0.02 0.64 ± 0.07
E20 − E10

19
Substituting the values for the ψ and Υ family and varying ν, we get

Fig. 4: Plot to determine ν

ψ family Υ family

ν 0.14 0.33

7.3. LOGARITHMIC POTENTIAL

Taking the mass differences for the first excited state and the ground state of the ψ and Υ
families we see that ΔMψ ≃ ΔMΥ.

We now look for a potential which gives a level spacing independent of the reduced mass.

( r0 )
r
We proved that the logarithmic potential V = C ln is unique in achieving this.

In the case of logarithmic potential we can express the mass of the quarkonium system as

Mnl = CEnl + E0(μ), (7.8)


where n and l are the principal and principal and orbital angular quantum number
respectively.

20
where E0 is dependent on μ and r0.

ψ
M1s = 3.096, E1s = 1.0443

ψ
M2s = 3.686, E2s = 1.8474
(7.9)
Υ
M1s = 9.46,

Υ
M2s = 10.02

3.096 = C(1.0443) + E0(μ)


3.686 = C(1.8474) + E0(μ), 0.59 = C(0.8031), C = 0.734 GeV (7.10)
Putting the value of C in the equation \\

E0(ψ) = 2.329 GeV (7.11)

E0(Υ) = 8.694 GeV (7.12)

C
From the virial theorem we had , for the logarithmic potential, that ⟨T ⟩ =
2
⟨ p 2⟩ 2
1 ⟨p ⟩
⟨β 2⟩ ∼ = m 2μ
⟨m 2⟩
⟨T ⟩
= m (7.13)
C
= 2m

which is roughly equal to 0.25. This is the reason we can describe the bound states of the
quark-antiquark pair non relativistically. And since ⟨β 2⟩ does not change for the excited
states the approximation is quite good for the excited states too. Also the non-relativistic
approximation becomes better for systems with increasing quark mass.

n\l 0 1 2 3 4

1 1.0443 1.643 2.015 2.286 2.499

2 1.8474 2.151 2.388 2.581 2.742

3 2.2897 2.491 2.663 2.811 2.911

Enl values for the logarithmic potential

21
7.4. CORNELL POTENTIAL

Since, the quarks have charges, the short range force arises from coulomb like interactions,
in which the coupling constant is weakly dependent on the mass of the system. The long
range force is like a restoring or confining force. Thus, the potential acquires the form,
α r
V(r) = − + 2, (7.14)
r a

The parameters α and a are chose in such a way that they can be used to describe the
charmonium and the upsilon family.

The values of α and a for which reproduced the the observed 1s and the 2s levels for the
charmonium and the upsilon family was achieved by the choices α = 0.506 and
a = 2.43GeV−1.

The Coulomb term becomes dominant with increase in quark mass which is according to the
eqn. 5.18. The quality of the non relativistic approximation increases with increased quark
mass.

The masses predicted from the logarithmic potential and the Coulomb potential are
tabulated below.

Charmonium Upsilon

Predicted mass Observerd Predicted mass Observerd


V≈ ln r V=-a/r + br mass V≈ ln r V=-a/r + br mass

1s - - 3.096 - - 9.46

2s - - 3.686 - - 10.016

3s 4.01 4.12 4.04 10.37 10.35 10.38±


0.04
- denotes the input data.
Predicted and observed masses for the two families

22
8. CONCLUSION
In this report a number of tools were presented which included the scaling techniques and
some other useful theorems like the Feynman Hellmann theorem and the virial theorem.
Using these the energy level spacings was generated using MATLAB. Also the behaviour of
the energy level differences were noted which varied with the reduced mass of the system
and also the coupling strength.

Some semiclassical results were taken which helped us in computing the effective power ν
of the power law potential V = λr ν for investigating the nature of the inter quark
interactions. Some aid was taken from the experimental results for computing ν.

We specifically looked at the two potentials namely the logarithmic potential and the
Cornell potential which generated level spacings in accordance with the experimental
findings. We were able to predict the masses of charmonium and upsilon family upto certain
accuracy using the two potentials. The logarithmic potential achieved energy level spacings
independent of quark masses. But it was unable to produce equal spacings for the two
families.

The non relativistic approach is no matter easy to work with but we should keep in mind
that we didn’t take the colour charge of the quarks into account. Also this method could
have been used because of the large mass of the two systems we worked on. The lighter
systems could not have been described in this way.

References
[1]  C. Quigg and Jonathan L. Rosner. Quantum mechanics with applications to
quarkonium. Physics Reports, 56(4):167 – 235, 1979. 


[2]  E. Eichten, K. Gottfried, T. Kinoshita, J. Kogut, K. D. Lane, and T. M. Yan. Spectrum of
charmed quark-antiquark bound states. Phys. Rev. Lett., 34:369–372, Feb 1975. 


[3] C. Quigg and Jonathan L. Rosner. Quarkonium Level Spacings. Phys. Lett., 71B:153–
157, 1977.

[4] Review of Particle Physics ,Particle Data Group (C. Patrignani (Bologna U. & INFN,
Bologna) et al.). 2016. 1808 pp.

23

You might also like