Previous Up Next
Creative Commons License
This work is licensed under a Creative Commons Attribution-NonCommercial-ShareAlike 4.0 International License.

Lecture 4 Harmonic Oscillator and Ladder Operators

Complex valued representations of the Heisenberg group (also known as Weyl or Heisenberg-Weyl group) provide a natural framework for quantum mechanics [139, 104]. This is the most fundamental example of the Kirillov orbit method, induced representations and geometrical quantisation technique [161, 160]. Following the presentation in Section 1.4 we will consider representations of the Heisenberg group which are induced by hypercomplex characters of its centre: complex (which correspond to the elliptic case), dual (parabolic) and double (hyperbolic).

To describe dynamics of a physical system we use a universal equation based on inner derivations (commutator) of the convolution algebra [177] [181]. The complex valued representations produce the standard framework for quantum mechanics with the Heisenberg dynamical equation [336].

The double number valued representations, with the hyperbolic unit є2=1, is a natural source of hyperbolic quantum mechanics developed for a while [143, 142, 154, 151, 155]. The universal dynamical equation employs hyperbolic commutator in this case. This can be seen as a Moyal bracket based on the hyperbolic sine function. The hyperbolic observables act as operators on a Krein space with an indefinite inner product. Such spaces are employed in study of PT-symmetric Hamiltonians and hyperbolic unit є2=1 naturally appear in this setup [121].

The representations with values in dual numbers provide a convenient description of the classical mechanics. For this we do not take any sort of semiclassical limit, rather the nilpotency of the parabolic unit (ε2=0) do the task. This removes the vicious necessity to consider the Planck constant tending to zero. The dynamical equation takes the Hamiltonian form. We also describe classical non-commutative representations of the Heisenberg group which acts in the first jet space.

Remark 1 It is worth to note that our technique is different from contraction technique in the theory of Lie groups [242, 119]. Indeed a contraction of the Heisenberg group n is the commutative Euclidean group 2n which does not recreate neither quantum nor classical mechanics.

The approach provides not only three different types of dynamics, it also generates the respective rules for addition of probabilities as well. For example, the quantum interference is the consequence of the same complex-valued structure, which directs the Heisenberg equation. The absence of an interference (a particle behaviour) in the classical mechanics is again the consequence the nilpotency of the parabolic unit. Double numbers creates the hyperbolic law of additions of probabilities, which was extensively investigates [154, 151]. There are still unresolved issues with positivity of the probabilistic interpretation in the hyperbolic case [143, 142].

Remark 2 It is commonly accepted since the Dirac’s paper [81] that the striking (or even the only) difference between quantum and classical mechanics is non-commutativity of observables in the first case. In particular the Heisenberg commutation relations (12) imply the uncertainty principle, the Heisenberg equation of motion and other quantum features. However, the entire book of Feynman on QED [97] does not contains any reference to non-commutativity. Moreover, our work shows that there is a non-commutative formulation of classical mechanics. Non-commutative representations of the Heisenberg group in dual numbers implies the Poisson dynamical equation and local addition of probabilities in Section4.5, which are completely classical.

This entirely dispels any illusive correlation between classical/quantum and commutative/non-commutative. Instead we show that quantum mechanics is fully determined by the properties of complex numbers. In Feynman’s exposition [97] complex numbers are presented by a clock, rotations of its arm encode multiplications by unimodular complex numbers. Moreover, there is no a presentation of quantum mechanics, which does not employ complex phases (numbers) in one or another form. Analogous parabolic and hyperbolic phases (or characters produced by associated hypercomplex numbers, see Section 3.1) lead to classical and hypercomplex mechanics respectively.

This section clarifies foundations of quantum and classical mechanics. We recovered the existence of three non-isomorphic models of mechanics from the representation theory. They were already derived in [143, 142] from translation invariant formulation, that is from the group theory as well. It also hinted that hyperbolic counterpart is (at least theoretically) as natural as classical and quantum mechanics are. The approach provides a framework for a description of aggregate system which have say both quantum and classical components. This can be used to model quantum computers with classical terminals [189].

Remarkably, simultaneously with the work [142] group-invariant axiomatics of geometry leaded R.I. Pimenov [284] to description of 3n Cayley–Klein constructions. The connection between group-invariant geometry and respective mechanics were explored in many works of N.A. Gromov, see for example [115, 116, 119]. They already highlighted the rôle of three types of hypercomplex units for the realisation of elliptic, parabolic and hyperbolic geometry and kinematic.

There is a further connection between representations of the Heisenberg group and hypercomplex numbers. The symplectomorphism of phase space are also automorphism of the Heisenberg group [104]*§ 1.2. We recall that the symplectic group [2] [104]*§ 1.2 is isomorphic to the group SL2(ℝ)  [240] [140] [251] and provides linear symplectomorphisms of the two-dimensional phase space. It has three types of non-iso­mor­phic one-dimensional continuous subgroups (3-5) with symplectic action on the phase space illustrated by Fig. 11.1. Hamiltonians, which produce those symplectomorphism, are of interest [338]*§ 3.8 [322] [323]. An analysis of those Hamiltonians from Section 3.3 by means of ladder operators recreates hypercomplex coefficients as well [196].

Harmonic oscillators, which we shall use as the main illustration here, are treated in most textbooks on quantum mechanics. This is efficiently done through creation/annihilation (ladder) operators, cf. § 3.3 and [109] [48]. The underlying structure is the representation theory of the Heisenberg and symplectic groups [240]*§ VI.2 [321]*§ 8.2 [139] [104]. As we will see, they are naturally connected with respective hypercomplex numbers. As a result we obtain further illustrations to the Similarity and Correspondence Principle 5.

We work with the simplest case of a particle with only one degree of freedom. Higher dimensions and the respective group of symplectomorphisms [2n] may require consideration of Clifford algebras [166] [69] [66] [121] [289].

4.1 p-Mechanic Formalism

Here we briefly outline a formalism [169, 294, 177, 53, 181], which allows to unify quantum and classical mechanics.

4.1.1 Convolutions (Observables) on ℍn and Commutator

Using a invariant measure dg=dsdxdy on ℍn we can define the convolution of two functions:

     
  (k1 * k2) (g)=
 


n
k1(g1)  k2(g1−1g) dg1  .  
(1)

This is a non-commutative operation, which is meaningful for functions from various spaces including L1(ℍn,dg), the Schwartz space S and many classes of distributions, which form algebras under convolutions. Convolutions on ℍn are used as observables in p-mechanic [169, 181].

A unitary representation of ℍn extends to L1(ℍn ,dg) by the formula:

 (k) = 
 


n
k(g)  (g) dg . (2)

This is also an algebra homomorphism of convolutions to linear operators.

For a dynamics of observables we need inner derivations Dk of the convolution algebra L1(ℍn), which are given by the commutator:

     
     Dk: f ↦ [k,f]=k*ff*k     (3)
 =
 


n
k(g1)
f(g1−1g)−f(gg1−1)
dg1 ,   f,kL1(ℍn).  
 

To describe dynamics of a time-dependent observable f(t,g) we use the universal equation, cf. [167, 169]:

Sḟ=[H,f], (4)

where S is the left-invariant vector field (10) generated by the centre of ℍn. The presence of operator S fixes the dimensionality of both sides of the equation (4) if the observable H (Hamiltonian) has the dimensionality of energy [181]*Rem 4.1. If we apply a right inverse of S to both sides of the equation (4) we obtain the equivalent equation

ḟ=Hf, (5)

based on the universal bracket k1k2=k1* k2k2* k1 [181].

Example 3 (Harmonic oscillator) Let H=1/2 (mk2 q2 + 1/mp2) be the Hamiltonian of a one-dimensional harmonic oscillator, where k is a constant frequency and m is a constant mass. Its p-mechanisation will be the second order differential operator on n [53]*§ 5.1:
     H=
1
2
 (mk2X2 + 
1
m
Y2),
where we dropped sub-indexes of vector fields (10) in one dimensional setting. We can express the commutator as a difference between the left and the right action of the vector fields:
     [H,f]=
1
2
 (mk2 ((Xr)2−(Xl)2) + 
1
m
((Yr)2−(Yl)2))f.
Thus the equation (4) becomes [53]*(5.2):
∂ 
∂ s
ḟ= 
∂ 
∂ s



mk2y
∂ x
1
m
x
∂ y



f.  (6)
Of course, the derivative ∂ /∂ s can be dropped from both sides of the equation and the general solution is found to be:
f(t;s,x,y)  =  f0


s, xcos(k t) + mkysin( kt),  −
x
mk
 sin(kt) + ycos(kt)


, (7)
where f0(s,x,y) is the initial value of an observable on n.
Example 4 (Unharmonic oscillator) We consider unharmonic oscillator with cubic potential, see [55] and references therein:
H=
mk2
2
q2+
λ
6
q3 + 
1
2m
p2. (8)
Due to the absence of non-commutative products p-mechanisation is straightforward:
    H=
mk2
2
  X2+
λ
6
X3 + 
1
m
Y2.
Similarly to the harmonic case the dynamic equation, after cancellation of ∂ /∂ s on both sides, becomes:
ḟ=     


mk2y
∂ x
+
λ
6



3y
2
∂ x2
  +
1
4
y3
2
∂ s2



1
m
x
∂ y



f.  (9)
Unfortunately, it cannot be solved analytically as easy as in the harmonic case.

4.1.2 States and Probability

Let an observable (k) (2) is defined by a kernel k(g) on the Heisenberg group and its representation at a Hilbert space H. A state on the convolution algebra is given by a vector vH. A simple calculation:

  ⟨ (k)v,v  ⟩H=
⟨ 
 


n
k(g) (g)vdg,v  ⟩H
 =
 


n
k(g) ⟨ (g)v,v  ⟩Hdg
 =
 


n
k(g) 
⟨ v,(g)v  ⟩H
dg

can be restated as:

  ⟨ (k)v,v  ⟩H=⟨ k,l  ⟩,    where   l(g)=⟨ v,(g)v  ⟩H.

Here the left-hand side contains the inner product on H, while the right-hand side uses a skew-linear pairing between functions on ℍn based on the Haar measure integration. In other words we obtain, cf. [53]*Thm. 3.11:

Proposition 5 A state defined by a vector vH coincides with the linear functional given by the wavelet transform
l(g)=⟨ v,(g)v  ⟩H (10)
of v used as the mother wavelet as well.

The addition of vectors in H implies the following operation on states:

     
  ⟨ v1+v2,(g)(v1+v2)  ⟩H= ⟨ v1,(g)v1  ⟩H +⟨ v2,(g)v2  ⟩H 
  
+⟨ v1,(g)v2  ⟩H + 
⟨ v1,(g−1)v2  ⟩H
 
(11)

The last expression can be conveniently rewritten for kernels of the functional as

l12=l1+l2+2 A
l1l2
(12)

for some real number A. This formula is behind the contextual law of addition of conditional probabilities [153] and will be illustrated below. Its physical interpretation is an interference, say, from two slits. Despite of a common belief, the mechanism of such interference can be both causal and local, see [179] [152].

4.2 Elliptic characters and Quantum Dynamics

In this section we consider the representation h of ℍn induced by the elliptic character χh(s)=eih s in complex numbers parametrised by h∈ℝ. We also use the convenient agreement h=2πℏ borrowed from physical literature.

4.2.1 Fock–Segal–Bargmann and Schrödinger Representations

The realisation of h by the left shifts (1) on L2h(ℍn) is rarely used in quantum mechanics. Instead two unitary equivalent forms are more common: the Schrödinger and Fock–Segal–Bargmann (FSB) representations.

The FSB representation can be obtained from the orbit method of Kirillov [160]. It allows spatially separate irreducible components of the left regular representation, each of them become located on the orbit of the co-adjoint representation, see [181]*§ 2.1 [160] for details, we only present a brief summary here.

We identify ℍn and its Lie algebra hn through the exponential map [159]*§ 6.4. The dual hn* of hn is presented by the Euclidean space ℝ2n+1 with coordinates (ℏ,q,p). The pairing hn* and hn given by

  ⟨ (s,x,y),(ℏ,q,p)  ⟩=ℏ s + q · x+p· y.

This pairing defines the Fourier transform ^: L2(ℍn)→ L2(hn*) given by [161]*§ 2.3:

φ(F)=
 


hn
 φ(expX)  e−2πi  ⟨ X,F  ⟩dX     where Xhn, Fhn*.  (13)

For a fixed ℏ the left regular representation (1) is mapped by the Fourier transform to the FSB type representation (13). The collection of points (ℏ,q,p)∈hn* for a fixed ℏ is naturally identified with the phase space of the system.

Remark 6 It is possible to identify the case of ℏ=0 with classical mechanics [181]. Indeed, a substitution of the zero value of into (13) produces the commutative representation:
0(s,x,y): f (q,p) ↦  e−2πi(qx+py) f
q, p
. (14)
It can be decomposed into the direct integral of one-dimensional representations parametrised by the points (q,p) of the phase space. The classical mechanics, including the Hamilton equation, can be recovered from those representations [181]. However the condition ℏ=0 (as well as the semiclassical limit ℏ→ 0) is not completely physical. Commutativity (and subsequent relative triviality) of those representation is the main reason why they are oftenly neglected. The commutativity can be outweighed by special arrangements, e.g. an antiderivative [181]*(4.1), but the procedure is not straightforward, see discussion in [184] [1] [188]. A direct approach using dual numbers will be shown below, cf. Rem. 18.

To recover the Schrödinger representation we use notations and technique of induced representations from § 3.2, see also [173]*Ex. 4.1. The subgroup H={(s,0,y) ∣  s∈ℝ, y∈ℝn}⊂ℍn defines the homogeneous space X=G/H, which coincides with ℝn as a manifold. The natural projection p:GX is p(s,x,y)=x and its left inverse s:XG can be as simple as s(x)=(0,x,0). For the map r:GH, r(s,x,y)=(sxy/2,0,y) we have the decomposition

  (s,x,y)=s(p(s,x,y))*r(s,x,y)=(0,x,0)*(s
1
2
xy,0,y).

For a character χh(s,0,y)=eih s of H the lifting Lχ: L2(G/H) → L2χ(G) is as follows:

  [Lχf](s,x,y)=χh(r(s,x,y))   f(p(s,x,y))=eih (sxy/2)f(x).  

Thus the representation χ(g)=P∘Λ (g)∘L becomes:

[χ(s′,x′,y′) f](x)=e−2πiℏ (s′+xy′−xy′/2)f(xx′).   (15)

After the Fourier transform xq we get the Schrödinger representation on the configuration space:

[χ(s′,x′,y′) f ](q)=e−2πiℏ (s′+xy′/2) −2πix′ q f(q+ℏ y′).   (16)

Note that this again turns into a commutative representation (multiplication by an unimodular function) if ℏ=0. To get the full set of commutative representations in this way we need to use the character χ(h,p)(s,0,y)=ei(ℏ+ py) in the above consideration.

4.2.2 Commutator and the Heisenberg Equation

The property (2) of F2χ(ℍn) implies that the restrictions of two operators χ (k1) and χ (k2) to this space are equal if

  
 


k1(s,x,y) χ(s)  ds = 
 


k2(s,x,y) χ(s) ds.

In other words, for a character χ(s)=eis the operator χ (k) depends only on

  ks(ℏ,x,y)=
 


k(s,x,y) e−2πi ℏ sds,

which is the partial Fourier transform s↦ ℏ of k(s,x,y). The restriction to F2χ(ℍn) of the composition formula for convolutions is [181]*(3.5):

(k′*k)_s=
 


2n
eih(xy′−yx′)/2  k′s(ℏ ,x′,y′)  ks(ℏ ,xx′,yy′) dxdy′.  (17)

Under the Schrödinger representation (16) the convolution (17) defines a rule for composition of two pseudo-differential operators (PDO) in the Weyl calculus [139] [104]*§ 2.3.

Consequently the representation (2) of commutator (3) depends only on its partial Fourier transform [181]*(3.6):

     
  [k′,k]_s=
   2 i  
 


2n
 sin(
h
2
(xy′−yx′))  
(18)
     ×  k′s(ℏ, x′, y′)  ks(ℏ, xx′, yy′) dxdy′.     

Under the Fourier transform (13) this commutator is exactly the Moyal bracket [341] for of k′ and k on the phase space.

For observables in the space F2χ(ℍn) the action of S is reduced to multiplication, e.g. for χ(s)=ei h s the action of S is multiplication by i h. Thus the equation (4) reduced to the space F2χ(ℍn) becomes the Heisenberg type equation [181]*(4.4):

ḟ=
1
ih
  [H,f]_s, (19)

based on the above bracket (18). The Schrödinger representation (16) transforms this equation to the original Heisenberg equation.

Example 7
  1. Under the Fourier transform (x,y)↦(q,p) the p-dynamic equation (6) of the harmonic oscillator becomes:
    ḟ=     


    mk2q
    ∂ p
    1
    m
    p
    ∂ q



    f.  (20)
    The same transform creates its solution out of (7).
  2. Since ∂/∂ s acts on F2χ(ℍn) as multiplication by i, the quantum representation of unharmonic dynamics equation (9) is:
    ḟ=     


    mk2q
    p
    +
    λ
    6



    3q2
    ∂ p
    2
    4
    3
    ∂ p3



    1
    m
    p
    ∂ q



    f.  (21)
    This is exactly the equation for the Wigner function obtained in [55]*(30).

4.2.3 Quantum Probabilities

For the elliptic character χh(s)=eih s we can use the Cauchy–Schwarz inequality to demonstrate that the real number A in the identity (12) is between −1 and 1. Thus we can put A=cosα for some angle (phase) α to get the formula for counting quantum probabilities, cf. [154]*(2):

l12=l1+l2+2 cosα  
l1l2
(22)
Remark 8 It is interesting to note that the both trigonometric functions are employed in quantum mechanics: sine is in the heart of the Moyal bracket (18) and cosine is responsible for the addition of probabilities (22). In the essence the commutator and probabilities took respectively the odd and even parts of the elliptic character eih s.
Example 9 Take a vector v(a,b)L2h(ℍn) defined by a Gaussian with mean value (a,b) in the phase space for a harmonic oscillator of the mass m and the frequency k:
v(a,b)(q,p)=exp


2π km
(qa)2
km
(pb)2


. (23)
A direct calculation shows:
  
⟨ v(a,b),ℏ(s,x,y)v(a′,b′)  ⟩=
4
exp


π i
2sℏ+x (a+a′)+y (b+b′)
 
  
 −
π
2 ℏ km
((ℏ x+bb′)2 +(bb′)2) −
π km
2ℏ
 ((ℏ y+a′−a)2 + (a′−a)2)


 =
4
exp


π i
2sℏ+x (a+a′)+y (b+b′)
 
  
  −
π
ℏ km
((bb′+
ℏ x
2
)2 +(
ℏ x
2
)2) −
π km
 ((aa′−
ℏ y
2
)2 + (
ℏ y
2
)2) 


Thus the kernel l(a,b)=⟨ v(a,b),ℏ(s,x,y)v(a,b) (10) for a state v(a,b) is:
     
  l(a,b)=
4
exp


2π i (sℏ+xa+yb)
 
πℏ
2 km
x2
π km ℏ
2ℏ
y2


 
(24)
An observable registering a particle at a point q=c of the configuration space is δ(qc). On the Heisenberg group this observable is given by the kernel:
Xc(s,x,y)=ei (sℏ+x  c)δ(y). (25)
The measurement of Xc on the state (23) (through the kernel (24)) predictably is:
  ⟨ Xc,l(a,b)  ⟩=
2k m
exp


2π km
(ca)2


.
Example 10 Now take two states v(0,b) and v(0,−b), where for the simplicity we assume the mean values of coordinates vanish in the both cases. Then the corresponding kernel (11) has the interference terms:
  li=⟨ v(0,b),ℏ(s,x,y)v(0,−b)  ⟩
 =
4
exp


2π isℏ −
π
2 ℏ km
((ℏ x+2b)2 +4b2) −
πℏ km
2
y2


.
The measurement of Xc (25) on this term contains the oscillating part:
  ⟨ Xc,li  ⟩=
2km
 exp


2π k m
c2
k m ℏ
b2+
i
cb


Therefore on the kernel l corresponding to the state v(0,b)+v(0,−b) the measurement is
  ⟨ Xc,l  ⟩=
2
2 k m
exp


2π km
c2





1+exp


k m ℏ
b2


cos


cb





.

(a)  (b)
Figure 4.1: Quantum probabilities: the blue (dashed) graph shows the addition of probabilities without interaction, the red (solid) graph present the quantum interference. Left picture shows the Gaussian state (23), the right—the rational state (26)

The presence of the cosine term in the last expression can generate an interference picture. In practise it does not happen for the minimal uncertainty state (23) which we are using here: it rapidly vanishes outside of the neighbourhood of zero, where oscillations of the cosine occurs, see Fig. 4.1(a).
Example 11 To see a traditional interference pattern one can use a state which is far from the minimal uncertainty. For example, we can consider the state:
u(a,b)(q,p)=
2
((qa)2+ℏ/ k m)((pb)2+ℏ km)
. (26)
To evaluate the observable Xc (25) on the state l(g)=⟨ u1,h(g)u2 (10) we use the following formula:
  ⟨ Xc,l  ⟩=
2
 


n
 û1(q, 2(qc)/ℏ)  
û2(q, 2(qc)/ℏ)
dq,
where ûi(q,x) denotes the partial Fourier transform px of ui(q,p). The formula is obtained by swapping order of integrations. The numerical evaluation of the state obtained by the addition u(0,b)+u(0,−b) is plotted on Fig. 4.1(b), the red curve shows the canonical interference pattern.

4.3 Ladder Operators and Harmonic Oscillator

Let be a representation of the Schrödinger group G=ℍ1Sp′(2) (17) in a space V. Consider the derived representation of the Lie algebra g [240]*§ VI.1 and denote X′=(X) for Xg. To see the structure of the representation we can decompose the space V into eigenspaces of the operator X′ for some Xg. The canonical example is the Taylor series in complex analysis.

We are going to consider three cases corresponding to three non-isomorphic subgroups (35) of [2] starting from the compact case. Let H=Z be a generator of the compact subgroup K. Corresponding symplectomorphisms (4) of the phase space are given by orthogonal rotations with matrices (

costsin t
−sintcost

). The Shale–Weil representation (16) coincides with the Hamiltonian of the harmonic oscillator in Schrödinger representation.

Since Sp′(2) is a two-fold cover the corresponding eigenspaces of a compact group Zvk=i k vk are parametrised by a half-integer k∈ℤ/2. Explicitly for a half-integer k eigenvectors are:

vk(q)=H
 
k+
1
2





q




e
π
q2
 
, (27)

where Hk is the Hermite polynomial [104]*§ 1.7 [93]*8.2(9).

From the point of view of quantum mechanics as well as the representation theory it is beneficial to introduce the ladder operators L± (15), known also as creation/annihilation in quantum mechanics [104]*p. 49 [48]. There are two ways to search for ladder operators: in (complexified) Lie algebras h1 and sp2. The later coincides with our consideration in Section 3.3 in the essence.

4.3.1 Ladder Operators from the Heisenberg Group

Assuming L+=aX′+bY′ we obtain from the relations (1819) and (15) the linear equations with unknown a and b:

  a+b,    −b+a.

The equations have a solution if and only if λ+2+1=0, and the raising/lowering operators are L±= X′∓iY′.

Remark 12 Here we have an interesting asymmetric response: due to the structure of the semidirect product 1Sp′(2) it is the symplectic group which acts on 1, not vice versa. However the Heisenberg group has a weak action in the opposite direction: it shifts eigenfunctions of [2].

In the Schrödinger representation (15) the ladder operators are

ℏ(L±)= 2πiq±iℏ 
d
dq
. (28)

The standard treatment of the harmonic oscillator in quantum mechanics, which can be found in many textbooks, e.g.  [104]*§ 1.7 [109]*§ 2.2.3, is as follows. The vector v−1/2(q)=e−π q2/ℏ is an eigenvector of Z′ with the eigenvalue −i/2. In addition v−1/2 is annihilated by L+. Thus the chain (16) terminates to the right and the complete set of eigenvectors of the harmonic oscillator Hamiltonian is presented by (L)k v−1/2 with k=0, 1, 2, ….

We can make a wavelet transform generated by the Heisenberg group with the mother wavelet v−1/2, and the image will be the Fock–Segal–Bargmann (FSB) space [139] [104]*§ 1.6. Since v−1/2 is the null solution of L+=X′−i Y′, then by Cor. 6 the image of the wavelet transform will be null-solutions of the corresponding linear combination of the Lie derivatives (10):

D=
Xr −i  Yr
=(∂x +iy)−πℏ(xi y), (29)

which turns out to be the Cauchy–Riemann equation on a weighted FSB-type space, see Section 5.4 below.

4.3.2 Symplectic Ladder Operators

We can also look for ladder operators within the Lie algebra sp2, see Subsection 4.3.1 and [194]*§ 8. Assuming L2+=aA′+bB′+cZ′ from the relations (13) and defining condition (15) we obtain the linear equations with unknown a, b and c:

  c=0,    2a+b,    −2b+a.

The equations have a solution if and only if λ+2+4=0, and the raising/lowering operators are L2±i A′+B′. In the Shale–Weil representation (16) they turn out to be:

L2±i


q
2
d
dq
+
1
4



i
d2
dq2
πiq2
2ℏ
=−
i
8πℏ



∓2π q+ℏ
d
dq



2



 
. (30)

Since this time λ+=2i the ladder operators L2± produce a shift on the diagram (16) twice bigger than the operators L± from the Heisenberg group. After all, this is not surprising since from the explicit representations (28) and (30) we get:

  L2±=−
i
8πℏ
(L±)2.

4.4 Hyperbolic Quantum Mechanics

Now we turn to double numbers also known as hyperbolic, split-complex, etc. numbers [339]*App. C [325] [157]. They form a two dimensional algebra ℝh spanned by 1 and є with the property є2=1. There are zero divisors:

  є±=
1
2
(1± j),    such that   є+ є=0    and   є±2±.

Thus double numbers algebraically isomorphic to two copies of ℝ spanned by є±. Being algebraically dull double numbers are nevertheless interesting as a homogeneous space [191, 194] and they are relevant in physics [151, 325, 326]. The combination of p-mechanical approach with hyperbolic quantum mechanics was already discussed in [53]*§ 6.

For the hyperbolic character χє h(s)=eє h s=coshh s +єsinhh s of ℝ one can define the hyperbolic Fourier-type transform:

  k(q)=
 


k(x) e−є qxdx.

It can be understood in the sense of distributions on the space dual to the set of analytic functions [155]*§ 3. Hyperbolic Fourier transform intertwines the derivative d/dx and multiplication by є q [155]*Prop. 1.

Example 13 For the Gaussian the hyperbolic Fourier transform is the ordinary function (note the sign difference!):
    
 


ex2/2e−є qxdx= 
  eq2/2.
However the opposite identity:
    
 


ex2/2e−є qxdx= 
  eq2/2
is true only in a suitable distributional sense. To this end we may note that ex2/2 and eq2/2 are null solutions to the differential operators d/dxx and d/dq+q respectively, which are intertwined (up to the factor є) by the hyperbolic Fourier transform. The above differential operators d/dxx and d/dq+q are images of the ladder operators (28) in the Lie algebra of the Heisenberg group. They are intertwining by the Fourier transform, since this is an automorphism of the Heisenberg group [138].

An elegant theory of hyperbolic Fourier transform may be achieved by a suitable adaptation of [138], which uses representation theory of the Heisenberg group.

4.4.1 Hyperbolic Representations of the Heisenberg Group

Consider the space Fhє(ℍn) of ℝh-valued functions on ℍn with the property:

f(s+s′,h,y)=eє hsf(s,x,y),     for all (s,x,y)∈ ℍn, s′∈ ℝ , (31)

and the square integrability condition (3). Then the hyperbolic representation is obtained by the restriction of the left shifts to Fhє(ℍn). To obtain an equivalent representation on the phase space we take ℝh-valued functional of the Lie algebra hn:

χ(h,q,p)j(s,x,y)=eє(hs +qx+ py) =cosh(hs +qx+ py) + єsinh(hs +qx+ py). (32)

The hyperbolic Fock–Segal–Bargmann type representation is intertwined with the left group action by means of the Fourier transform (13) with the hyperbolic functional (32). Explicitly this representation is:

ℏ(s,x,y): f (q,p) ↦   e−є(hs+qx+py) f


q
h
2
y, p+
h
2
x


. (33)

For a hyperbolic Schrödinger type representation we again use the scheme described in § 3.2. Similarly to the elliptic case one obtains the formula, resembling (15):

[єχ(s′,x′,y′) f](x)=e−єh (s′+xy′−xy′/2)f(xx′). (34)

Application of the hyperbolic Fourier transform produces a Schrödinger type representation on the configuration space, cf. (16):

  [єχ(s′,x′,y′) f ](q)=e−єh (s′+xy′/2) −є x′ q f(q+hy′).  

The extension of this representation to kernels according to (2) generates hyperbolic pseudodifferential operators introduced in [155]*(3.4).

4.4.2 Hyperbolic Dynamics

Similarly to the elliptic (quantum) case we consider a convolution of two kernels on ℍn restricted to Fhє(ℍn). The composition law becomes, cf. (17):

(k′*k)_s=
 


2n
e є h(xy′−yx′)  k′s(h ,x′,y′)  ks(h ,xx′,yy′) dxdy′.  (35)

This is close to the calculus of hyperbolic PDO obtained in [155]*Thm. 2. Respectively for the commutator of two convolutions we get, cf. (18):

[k′,k]_s= 
 


2n
 sinh(h (xy′−yx′))  k′s(h ,x′,y′)  ks(h ,xx′,yy′) dxdy′.  (36)

This the hyperbolic version of the Moyal bracket, cf. [155]*p. 849, which generates the corresponding image of the dynamic equation (4).

Example 14
  1. For a quadratic Hamiltonian, e.g. harmonic oscillator from Example 3, the hyperbolic equation and respective dynamics is identical to quantum considered before.
  2. Since ∂/∂ s acts on F2є(ℍn) as multiplication by є h and є2=1, the hyperbolic image of the unharmonic equation (9) becomes:
          ḟ=     


    m  k2q
    p
    +
    λ
    6



    3q2
    ∂ p
    +
    2
    4
    3
    ∂ p3



    1
    m
    p
    ∂ q



    f. 
    The difference with quantum mechanical equation (21) is in the sign of the cubic derivative.

4.4.3 Hyperbolic Probabilities


(a)  (b)
Figure 4.2: Hyperbolic probabilities: the blue (dashed) graph shows the addition of probabilities without interaction, the red (solid) graph present the quantum interference. Left picture shows the Gaussian state (23), with the same distribution as in quantum mechanics, cf. Fig. 4.1(a). The right picture shows the rational state (26), note the absence of interference oscillations in comparison with the quantum state on Fig. 4.1(b).

To calculate probability distribution generated by a hyperbolic state we are using the general procedure from Section 4.1.2. The main differences with the quantum case are as follows:

  1. The real number A in the expression (12) for the addition of probabilities is bigger than 1 in absolute value. Thus it can be associated with the hyperbolic cosine coshα , cf. Rem. 8, for certain phase α∈ℝ [155].
  2. The nature of hyperbolic interference on two slits is affected by the fact that eє h s is not periodic and the hyperbolic exponent eє t and cosine cosht do not oscillate. It is worth to notice that for Gaussian states the hyperbolic interference is exactly the same as quantum one, cf. Figs. 4.1(a) and 4.2(a). This is similar to coincidence of quantum and hyperbolic dynamics of harmonic oscillator.

    The contrast between two types of interference is prominent for the rational state (26), which is far from the minimal uncertainty, see the different patterns on Figs. 4.1(b) and 4.2(b).

4.4.4 Ladder Operators for the Hyperbolic Subgroup

Consider the case of the Hamiltonian H=2B, which is a repulsive (hyperbolic) harmonic oscillator [338]*§ 3.8. The corresponding one-dimensional subgroup of symplectomorphisms produces hyperbolic rotations of the phase space, see Fig. 11.1. The eigenvectors vµ of the operator

  SWℏ(2B)vν=−i


d2
dq2
+
π q2



vν=iν vν, 

are Weber–Hermite (or parabolic cylinder) functions vν=Dν−1/2(±2ei π/4π/ℏ q), see [93]*§ 8.2 [312] for fundamentals of Weber–Hermite functions and [322] for further illustrations and applications in optics.

The corresponding one-parameter group is not compact and the eigenvalues of the operator 2B′ are not restricted by any integrality condition, but the raising/lowering operators are still important [140]*§ II.1 [251]*§ 1.1. We again seek solutions in two subalgebras h1 and sp2 separately. However the additional options will be provided by a choice of the number system: either complex or double.

Example 15 (Complex Ladder Operators)

Assuming Lh+=aX′+bY from the commutators (1819) we obtain the linear equations:

a+b,    −b+a. (37)

The equations have a solution if and only if λ+2−1=0. Taking the real roots λ=±1 we obtain that the raising/lowering operators are Lh±=X′∓Y. In the Schrödinger representation (15) the ladder operators are

Lh±= 2πiq± ℏ 
d
dq
. (38)

The null solutions v±1/2(q)=e±πi/ℏ q2 to operators ℏ(L±) are also eigenvectors of the Hamiltonian SWℏ(2B) with the eigenvalue ±1/2. However the important distinction from the elliptic case is, that they are not square-integrable on the real line anymore.

We can also look for ladder operators within the sp2, that is in the form L2h+=aA′+bB′+cZ for the commutator [2B′,Lh+]=λ Lh+, see § ??. Within complex numbers we get only the values λ=± 2 with the ladder operators L2h±=±2A′+Z′/2, see [140]*§ II.1 [251]*§ 1.1. Each indecomposable h1- or sp2-module is formed by a one-dimensional chain of eigenvalues with a transitive action of ladder operators Lh± or L2h± respectively. And we again have a quadratic relation between the ladder operators:

  L2h±=
i
4πℏ
(Lh±)2.

4.4.5 Double Ladder Operators

There are extra possibilities in in the context of hyperbolic quantum mechanics [154] [151] [155]. Here we use the representation of ℍ1 induced by a hyperbolic character eє h t=cosh(h t)+єsinh(h t), see [199]*(4.5), and obtain the hyperbolic representation of ℍ1, cf. (16):

h(s′,x′,y′) f ](q)=eєh (s′−xy′/2) +є x′ q f(qhy′).   (39)

The corresponding derived representation is

єh(X)=є q,   єh(Y)=−h
d
dq
,    єh(S)=єhI. (40)

Then the associated Shale–Weil derived representation of sp2 in the Schwartz space S(ℝ) is, cf. (16):

SWh(A) =−
q
2
d
dq
1
4
,  SWh(B)=
єh
4
d2
dq2
є q2
4h
,  SWh(Z)=−
єh
2
d2
dq2
є q2
2h
. (41)

Note that SWh(B) now generates a usual harmonic oscillator, not the repulsive one like SWℏ(B) in (16). However the expressions in the quadratic algebra are still the same (up to a factor), cf. (1719):

     
      SWh(A)=
є
2h
h(Xh(Y) −
1
2
єh(S))
(42)
 =
є
4h
h(Xh(Y) +єh(Yh(X)),  
 
    SWh(B)=
є
4h
h(X)2−єh(Y)2), 
(43)
    SWh(Z)=
є
2h
h(X)2h(Y)2). 
(44)

This is due to the Principle 5 of similarity and correspondence: we can swap operators Z and B with simultaneous replacement of hypercomplex units i and є.

The eigenspace of the operator 2SWh(B) with an eigenvalue є ν are spanned by the Weber–Hermite functions D−ν−1/2(±√2/hx), see [93]*§ 8.2. Functions Dν are generalisations of the Hermit functions (27).

The compatibility condition for a ladder operator within the Lie algebra h1 will be (37) as before, since it depends only on the commutators (1819). Thus we still have the set of ladder operators corresponding to values λ=±1:

  Lh±=X′∓Y′=є q±h
d
dq
.

Admitting double numbers we have an extra way to satisfy λ2=1 in (37) with values λ=±є. Then there is an additional pair of hyperbolic ladder operators, which are identical (up to factors) to (28):

  Lє±=X′∓єY′=є q±єh
d
dq
.

Pairs Lh± and Lє± shift eigenvectors in the “orthogonal” directions changing their eigenvalues by ±1 and ±є. Therefore an indecomposable sp2-module can be para­metrised by a two-dimensional lattice of eigenvalues in double numbers, see Fig. 3.2.

The following functions

  v
±h
 
1
2
(q)
=
e∓є q2/(2h)=cosh
q2
2h
∓ єsinh
q2
2h
,
  v
±є
 
1
2
(q)
=e∓  q2/(2h)

are null solutions to the operators Lh± and Lє± respectively. They are also eigenvectors of 2SWh(B) with eigenvalues ∓є/2 and ∓1/2 respectively. If these functions are used as mother wavelets for the wavelet transforms generated by the Heisenberg group, then the image space will consist of the null-solutions of the following differential operators, see Cor. 6:

Dh=
Xr − Yr
=(∂x −∂y)+
h
2
(x+y),    Dє=
Xr − є Yr
=(∂x +є∂y)−
h
2
(x−є y),

for v1/2±h and v1/2±є respectively. This is again in line with the classical result (29). However annihilation of the eigenvector by a ladder operator does not mean that the part of the 2D-lattice becomes void since it can be reached via alternative routes on this lattice. Instead of multiplication by a zero, as it happens in the elliptic case, a half-plane of eigenvalues will be multiplied by the divisors of zero 1±є.

We can also search ladder operators within the algebra sp2 and admitting double numbers we will again find two sets of them, cf. § ??:

  L2h±=
±A′+Z′/2 = ∓
q
2
d
dq
1
4
− 
єh
4
d2
dq2
є q2
4h
=−
є
4h
(Lh±)2,
  L±=
±єA′+Z′/2=   ∓
є q
2
d
dq
є
4
єh
4
d2
dq2
є q2
4h
=−
є
4h
(Lє±)2.

Again the operators L2h± and L2h± produce double shifts in the orthogonal directions on the same two-dimensional lattice in Fig. 3.2.

4.5 Parabolic (Classical) Representations on the Phase Space

After the previous two cases it is natural to link classical mechanics with dual numbers generated by the parabolic unit ε2=0. Connection of the parabolic unit ε with the Galilean group of symmetries of classical mechanics is around for a while [339]*App. C.

However the nilpotency of the parabolic unit ε make it difficult if we will work with dual number valued functions only. To overcome this issue we consider a commutative real algebra C spanned by 1, i, ε and iε with identities i2=−1 and ε2=0. A seminorm on C is defined as follows:

  
a+bi+cε+diε 
2=a2+b2.

4.5.1 Classical Non-Commutative Representations

We wish to build a representation of the Heisenberg group which will be a classical analog of the Fock–Segal–Barg­mann representation (13). To this end we introduce the space Fhε(ℍn) of C-valued functions on ℍn with the property:

f(s+s′,h,y)=eε hsf(s,x,y),     for all (s,x,y)∈ ℍn, s′∈ ℝ , (45)

and the square integrability condition (3). It is invariant under the left shifts and we restrict the left group action to Fhε(ℍn).

An equivalent form of the induced representation acts on Fhε(ℝ2n), where ℝ2n) is the homogeneous space of ℍn over its centre. The Fourier transform (x,y)↦(q,p) intertwines the last representation with the following action on C-valued functions on the phase space:

εh(s,x,y): f(q,p) ↦ e−2πi(xq+yp)(f(q,p) +εh(sf(q,p) +
y
i
fq(q,p)−
x
i
fp(q,p))). (46)
Remark 16 Comparing the traditional infinite-dimensional (13) and one-dimen­sional (14) representations of n we can note that the properties of the representation (46) are a non-trivial mixture of the former:
  1. The action (46) is non-commutative, similarly to the quantum representation (13) and unlike the classical one (14). This non-commutativity will produce the Hamilton equations below in a way very similar to Heisenberg equation, see Rem. 18.
  2. The representation (46) does not change the support of a function f on the phase space, similarly to the classical representation (14) and unlike the quantum one (13). Such a localised action will be responsible later for an absence of an interference in classical probabilities.
  3. The parabolic representation (46) can not be derived from either the elliptic (13) or hyperbolic (33) by the plain substitution h=0.

We may also write a classical Schrödinger type representation. According to § 3.2 we get a representation formally very similar to the elliptic (15) and hyperbolic versions (34):

     
      [εχ(s′,x′,y′) f](x)=e−εh (s′+xy′−xy′/2)f(xx′)(47)
 =
(1−εh (s′+xy′−
1
2
xy′)) f(xx′).  
 

However due to nilpotency of ε the (complex) Fourier transform xq produces a different formula for parabolic Schrödinger type representation in the configuration space, cf. (16) and (39):

    [εχ(s′,x′,y′) f](q)= eix′ q





1−εh (s′−
1
2
xy′)


f(q) +
εhy
i
f′(q)


.

This representation shares all properties mentioned in Rem. 16 as well.

4.5.2 Hamilton Equation

The identity e ε te −ε t= 2ε t can be interpreted as a parabolic version of the sine function, while the parabolic cosine is identically equal to one, cf. § 3.1 and [129, 190]. From this we obtain the parabolic version of the commutator (18):

  [k′,k]_s(ε h, x,y)=
  ε h
 


2n
(xy′−yx′) 
  ×  k′s(ε h,x′,y′)    ks(ε h,xx′,yy′) dxdy′,   

for the partial parabolic Fourier-type transform ks of the kernels. Thus the parabolic representation of the dynamical equation (4) becomes:

εh
dfs
dt
h,x,y;t)= ε h
 


2n
(xy′−yx′)  Ĥs(ε h,x′,y′)    fs(ε h,xx′,yy′;t) dxdy′, (47)

Although there is no possibility to divide by ε (since it is a zero divisor) we can obviously eliminate ε h from the both sides if the rest of the expressions are real. Moreover this can be done “in advance” through a kind of the antiderivative operator considered in [181]*(4.1). This will prevent “imaginary parts” of the remaining expressions (which contain the factor ε) from vanishing.

Remark 17 It is noteworthy that the Planck constant completely disappeared from the dynamical equation. Thus the only prediction about it following from our construction is h≠ 0, which was confirmed by experiments, of course.

Using the duality between the Lie algebra of ℍn and the phase space we can find an adjoint equation for observables on the phase space. To this end we apply the usual Fourier transform (x,y)↦(q,p). It turn to be the Hamilton equation [181]*(4.7). However the transition to the phase space is more a custom rather than a necessity and in many cases we can efficiently work on the Heisenberg group itself.

Remark 18 It is noteworthy, that the non-commutative representation (46) allows to obtain the Hamilton equation directly from the commutator h(k1),εh(k2)]. Indeed its straightforward evaluation will produce exactly the above expression. By contrast, such a commutator for the commutative representation (14) is zero and to obtain the Hamilton equation we have to work with an additional tools, e.g. an anti-derivative [181]*(4.1).
Example 19
  1. For the harmonic oscillator in Example 3 the equation (47) again reduces to the form (6) with the solution given by (7). The adjoint equation of the harmonic oscillator on the phase space is not different from the quantum written in Example 7(1). This is true for any Hamiltonian of at most quadratic order.
  2. For non-quadratic Hamiltonians classical and quantum dynamics are different, of course. For example, the cubic term of s in the equation (9) will generate the factor ε3=0 and thus vanish. Thus the equation (47) of the unharmonic oscillator on n becomes:
          ḟ=     


    m  k2y
    ∂ x
    +
    λ y
    2
    2
    ∂ x2
      −
    1
    m
    x
    ∂ y



    f. 
    The adjoint equation on the phase space is:
          ḟ=     





    m  k2 q+
    λ
    2
    q2


    ∂ p
     −
    1
    m
      p
    ∂ q



    f. 
    The last equation is the classical Hamilton equation generated by the cubic potential (8). Qualitative analysis of its dynamics can be found in many textbooks [11]*§ 4.C, Pic. 12 [278]*§ 4.4.
Remark 20 We have obtained the Poisson bracket from the commutator of convolutions on n without any quasiclassical limit h→ 0. This has a common source with the deduction of main calculus theorems in [58] based on dual numbers. As explained in [191]*Rem. 6.9 this is due to the similarity between the parabolic unit ε and the infinitesimal number used in non-standard analysis [73]. In other words, we never need to take care about terms of order O(h2) because they will be wiped out by ε2=0.

An alternative derivation of classical dynamics from the Heisenberg group is given in the recent paper [247].

4.5.3 Classical probabilities

It is worth to notice that dual numbers are not only helpful in reproducing classical Hamiltonian dynamics, they also provide the classic rule for addition of probabilities. We use the same formula (10) to calculate kernels of the states. The important difference now that the representation (46) does not change the support of functions. Thus if we calculate the correlation term ⟨ v1,(g)v2 ⟩ in (11), then it will be zero for every two vectors v1 and v2 which have disjoint supports in the phase space. Thus no interference similar to quantum or hyperbolic cases (Section 4.2.3) is possible.

4.5.4 Ladder Operator for the Nilpotent Subgroup

Finally we look for ladder operators for the Hamiltonian B′+Z′/2 or, equivalently, −B′+Z′/2. It can be identified with a free particle [338]*§ 3.8.

We can look for ladder operators in the representation (1516) within the Lie algebra h1 in the form Lε±=aX′+bY′. This is possible if and only if

b=λ a,  0=λ b. (48)

The compatibility condition λ2=0 implies λ=0 within complex numbers. However such a “ladder” operator produces only the zero shift on the eigenvectors, cf. (??).

Another possibility appears if we consider the representation of the Heisenberg group induced by dual-valued characters. On the configuration space such a representation is [199]*(4.11):

[εχ(s,x,y) f](q)= eixq





1−εh (s
1
2
xy)


f(q) +
εhy
i
f′(q)


. (49)

The corresponding derived representation of h1 is

ph(X)=2πiq,   ph(Y)=
εh
2π i
d
dq
,    ph(S)=−εhI. (50)

However the Shale–Weil extension generated by this representation is inconvenient. It is better to consider the FSB–type parabolic representation (46) on the phase space induced by the same dual-valued character. Then the derived representation of h1 is:

ph(X)=−2πiq
εh
i
p,   ph(Y)=−2πip+
εh
i
q,    ph(S)=εhI. (51)

An advantage of the FSB representation is that the derived form of the parabolic Shale–Weil representation coincides with the elliptic one (22).

Eigenfunctions with the eigenvalue µ of the parabolic Hamiltonian B′+Z′/2=qp have the form

vµ(q,p)=eµ p/qf(q),  with an arbitrary function f(q). (52)

The linear equations defining the corresponding ladder operator Lε±=aX′+bY′ in the algebra h1 are (48). The compatibility condition λ2=0 implies λ=0 within complex numbers again. Admitting dual numbers we have additional values λ=±ελ1 with λ1∈ℂ with the corresponding ladder operators

  Lε±=X′∓ελ1Y′= −2πiq
εh
i
p± 2πελ1ip=  −2πiq+   εi( ± 2πλ1p+
h
p).

For the eigenvalue µ=µ0+εµ1 with µ0, µ1∈ℂ the eigenfunction (52) can be rewritten as:

vµ(q,p)=eµ  p/qf(q)= eµ0  p/q


1+εµ1
p
q



f(q) (53)

due to the nilpotency of ε. Then the ladder action of Lε± is µ0+εµ1↦ µ0+ε(µ1± λ1). Therefore these operators are suitable for building sp2-modules with a one-dimensional chain of eigenvalues.

Finally, consider the ladder operator for the same element B+Z/2 within the Lie algebra sp2, cf. § ??. There is the only operator Lp±=B′+Z′/2 corresponding to complex coefficients, which does not affect the eigenvalues. However the dual numbers lead to the operators

  Lε±=± ελ2A′+B′+Z′/2 = ±
ελ2
2

qqpp
+qp,     λ2∈ℂ. 

These operator act on eigenvalues in a non-trivial way.

4.5.5 Similarity and Correspondence

We wish to summarise our findings. Firstly, the appearance of hypercomplex numbers in ladder operators for h1 follows exactly the same pattern as was already noted for sp2, see Rem. 9:

In the spirit of the Similarity and Correspondence Principle 5 we have the following extension of Prop. 10:

Proposition 21 Let a vector Hsp2 generates the subgroup K, N or , that is H=Z, B+Z/2, or 2B respectively. Let ι be the respective hypercomplex unit. Then the ladder operators L± satisfying to the commutation relation:
    [H,L2±]=±ι L±
are given by:
  1. Within the Lie algebra h1: L±=X′∓ι Y′.
  2. Within the Lie algebra sp2: L2±=±ι A′ +E. Here Esp2 is a linear combination of B and Z with the properties:
    • E=[A,H].
    • H=[A,E].
    • Killings form K(H,E) [159]*§ 6.2 vanishes.
    Any of the above properties defines the vector Espan{B,Z} up to a real constant factor.

It is worth continuing this investigation and describing in details hyperbolic and parabolic versions of FSB spaces.

site search by freefind advanced

Last modified: October 28, 2024.
Previous Up Next