The Mehler kernel is a complex-valued function found to be the propagator of the quantum harmonic oscillator .
Mehler (1866 ) defined a function[ 1]
E
(
x
,
y
)
=
1
1
−
ρ
2
exp
(
−
ρ
2
(
x
2
+
y
2
)
−
2
ρ
x
y
(
1
−
ρ
2
)
)
,
{\displaystyle E(x,y)={\frac {1}{\sqrt {1-\rho ^{2}}}}\exp \left(-{\frac {\rho ^{2}(x^{2}+y^{2})-2\rho xy}{(1-\rho ^{2})}}\right)~,}
and showed, in modernized notation,[ 2] that it can be expanded in terms of Hermite polynomials H (.) based on weight function exp(−x ²) as
E
(
x
,
y
)
=
∑
n
=
0
∞
(
ρ
/
2
)
n
n
!
H
n
(
x
)
H
n
(
y
)
.
{\displaystyle E(x,y)=\sum _{n=0}^{\infty }{\frac {(\rho /2)^{n}}{n!}}~{\mathit {H}}_{n}(x){\mathit {H}}_{n}(y)~.}
This result is useful, in modified form, in quantum physics, probability theory, and harmonic analysis.
Physics version
In physics, the fundamental solution , (Green's function ), or propagator of the Hamiltonian for the quantum harmonic oscillator is called the Mehler kernel . It provides the fundamental solution [ 3] φ (x ,t ) to
∂
φ
∂
t
=
∂
2
φ
∂
x
2
−
x
2
φ
≡
D
x
φ
.
{\displaystyle {\frac {\partial \varphi }{\partial t}}={\frac {\partial ^{2}\varphi }{\partial x^{2}}}-x^{2}\varphi \equiv D_{x}\varphi ~.}
The orthonormal eigenfunctions of the operator D are the Hermite functions ,
ψ
n
=
H
n
(
x
)
exp
(
−
x
2
/
2
)
2
n
n
!
π
,
{\displaystyle \psi _{n}={\frac {H_{n}(x)\exp(-x^{2}/2)}{\sqrt {2^{n}n!{\sqrt {\pi }}}}},}
with corresponding eigenvalues (-2n -1), furnishing particular solutions
φ
n
(
x
,
t
)
=
e
−
(
2
n
+
1
)
t
H
n
(
x
)
exp
(
−
x
2
/
2
)
.
{\displaystyle \varphi _{n}(x,t)=e^{-(2n+1)t}~H_{n}(x)\exp(-x^{2}/2)~.}
The general solution is then a linear combination of these; when fitted to the initial condition φ (x ,0) , the general solution reduces to
φ
(
x
,
t
)
=
∫
K
(
x
,
y
;
t
)
φ
(
y
,
0
)
d
y
,
{\displaystyle \varphi (x,t)=\int K(x,y;t)\varphi (y,0)dy~,}
where the kernel K has the separable representation
K
(
x
,
y
;
t
)
≡
∑
n
≥
0
e
−
(
2
n
+
1
)
t
π
2
n
n
!
H
n
(
x
)
H
n
(
y
)
exp
(
−
(
x
2
+
y
2
)
/
2
)
.
{\displaystyle K(x,y;t)\equiv \sum _{n\geq 0}{\frac {e^{-(2n+1)t}}{{\sqrt {\pi }}2^{n}n!}}~H_{n}(x)H_{n}(y)\exp(-(x^{2}+y^{2})/2)~.}
Utilizing Mehler's formula then yields
∑
n
≥
0
(
ρ
/
2
)
n
n
!
H
n
(
x
)
H
n
(
y
)
exp
(
−
(
x
2
+
y
2
)
/
2
)
=
1
(
1
−
ρ
2
)
exp
(
4
x
y
ρ
−
(
1
+
ρ
2
)
(
x
2
+
y
2
)
2
(
1
−
ρ
2
)
)
.
{\displaystyle {\sum _{n\geq 0}{\frac {(\rho /2)^{n}}{n!}}H_{n}(x)H_{n}(y)\exp(-(x^{2}+y^{2})/2)={1 \over {\sqrt {(1-\rho ^{2})}}}\exp \left({4xy\rho -(1+\rho ^{2})(x^{2}+y^{2}) \over 2(1-\rho ^{2})}\right)}~.}
On substituting this in the expression for K with the value exp(−2t ) for ρ , Mehler's kernel finally reads
K
(
x
,
y
;
t
)
=
1
2
π
sinh
(
2
t
)
exp
(
−
coth
(
2
t
)
(
x
2
+
y
2
)
/
2
+
csch
(
2
t
)
x
y
)
.
{\displaystyle K(x,y;t)={\frac {1}{\sqrt {2\pi \sinh(2t)}}}~\exp \left(-\coth(2t)~(x^{2}+y^{2})/2+\operatorname {csch} (2t)~xy\right).}
When t = 0, variables x and y coincide, resulting in the limiting formula necessary by the initial condition,
K
(
x
,
y
;
0
)
=
δ
(
x
−
y
)
.
{\displaystyle K(x,y;0)=\delta (x-y)~.}
As a fundamental solution, the kernel is additive,
∫
d
y
K
(
x
,
y
;
t
)
K
(
y
,
z
;
t
′
)
=
K
(
x
,
z
;
t
+
t
′
)
.
{\displaystyle \int dyK(x,y;t)K(y,z;t')=K(x,z;t+t')~.}
This is further related to the symplectic rotation structure of the kernel K .[ 4]
When using the usual physics conventions of defining the quantum harmonic oscillator instead via
i
∂
φ
∂
t
=
1
2
(
−
∂
2
∂
x
2
+
x
2
)
φ
≡
H
φ
,
{\displaystyle i{\frac {\partial \varphi }{\partial t}}={\frac {1}{2}}\left(-{\frac {\partial ^{2}}{\partial x^{2}}}+x^{2}\right)\varphi \equiv H\varphi ,}
and assuming natural length and energy scales , then the Mehler kernel becomes the Feynman propagator
K
H
{\displaystyle K_{H}}
which reads
⟨
x
∣
exp
(
−
i
t
H
)
∣
y
⟩
≡
K
H
(
x
,
y
;
t
)
=
1
2
π
i
sin
t
exp
(
i
2
sin
t
(
(
x
2
+
y
2
)
cos
t
−
2
x
y
)
)
,
t
<
π
,
{\displaystyle \langle x\mid \exp(-itH)\mid y\rangle \equiv K_{H}(x,y;t)={\frac {1}{\sqrt {2\pi i\sin t}}}\exp \left({\frac {i}{2\sin t}}\left((x^{2}+y^{2})\cos t-2xy\right)\right),\quad t<\pi ,}
i.e.
K
H
(
x
,
y
;
t
)
=
K
(
x
,
y
;
i
t
/
2
)
.
{\displaystyle K_{H}(x,y;t)=K(x,y;it/2).}
When
t
>
π
{\displaystyle t>\pi }
the
i
sin
t
{\displaystyle i\sin t}
in the inverse square-root should be replaced by
|
sin
t
|
{\displaystyle |\sin t|}
and
K
H
{\displaystyle K_{H}}
should be
multiplied by an extra Maslov phase factor [ 5]
exp
(
i
θ
M
a
s
l
o
v
)
=
exp
(
−
i
π
2
(
1
2
+
⌊
t
π
⌋
)
)
.
{\displaystyle \exp \left(i\theta _{\rm {Maslov}}\right)=\exp \left(-i{\frac {\pi }{2}}\left({\frac {1}{2}}+\left\lfloor {\frac {t}{\pi }}\right\rfloor \right)\right).}
When
t
=
π
/
2
{\displaystyle t=\pi /2}
the general solution is proportional to the Fourier transform
F
{\displaystyle {\mathcal {F}}}
of the initial conditions
φ
0
(
y
)
≡
φ
(
y
,
0
)
{\displaystyle \varphi _{0}(y)\equiv \varphi (y,0)}
since
φ
(
x
,
t
=
π
/
2
)
=
∫
K
H
(
x
,
y
;
π
/
2
)
φ
(
y
,
0
)
d
y
=
1
2
π
i
∫
exp
(
−
i
x
y
)
φ
(
y
,
0
)
d
y
=
exp
(
−
i
π
/
4
)
F
[
φ
0
]
(
x
)
,
{\displaystyle \varphi (x,t=\pi /2)=\int K_{H}(x,y;\pi /2)\varphi (y,0)dy={\frac {1}{\sqrt {2\pi i}}}\int \exp(-ixy)\varphi (y,0)dy=\exp(-i\pi /4){\mathcal {F}}[\varphi _{0}](x)~,}
and the exact Fourier transform is thus obtained from the quantum harmonic oscillator's number operator written as[ 6]
N
≡
1
2
(
x
−
∂
∂
x
)
(
x
+
∂
∂
x
)
=
H
−
1
2
=
1
2
(
−
∂
2
∂
x
2
+
x
2
−
1
)
{\displaystyle N\equiv {\frac {1}{2}}\left(x-{\frac {\partial }{\partial x}}\right)\left(x+{\frac {\partial }{\partial x}}\right)=H-{\frac {1}{2}}={\frac {1}{2}}\left(-{\frac {\partial ^{2}}{\partial x^{2}}}+x^{2}-1\right)~}
since the resulting kernel
⟨
x
∣
exp
(
−
i
t
N
)
∣
y
⟩
≡
K
N
(
x
,
y
;
t
)
=
exp
(
i
t
/
2
)
K
H
(
x
,
y
;
t
)
=
exp
(
i
t
/
2
)
K
(
x
,
y
;
i
t
/
2
)
{\displaystyle \langle x\mid \exp(-itN)\mid y\rangle \equiv K_{N}(x,y;t)=\exp(it/2)K_{H}(x,y;t)=\exp(it/2)K(x,y;it/2)}
also compensates for the phase factor still arising in
K
H
{\displaystyle K_{H}}
and
K
{\displaystyle K}
, i.e.
φ
(
x
,
t
=
π
/
2
)
=
∫
K
N
(
x
,
y
;
π
/
2
)
φ
(
y
,
0
)
d
y
=
F
[
φ
0
]
(
x
)
,
{\displaystyle \varphi (x,t=\pi /2)=\int K_{N}(x,y;\pi /2)\varphi (y,0)dy={\mathcal {F}}[\varphi _{0}](x)~,}
which shows that the number operator can be interpreted via the Mehler kernel as the generator of fractional Fourier transforms for arbitrary values of t , and of the conventional Fourier transform
F
{\displaystyle {\mathcal {F}}}
for the particular value
t
=
π
/
2
{\displaystyle t=\pi /2}
, with the Mehler kernel providing an active transform , while the corresponding passive transform is already embedded in the basis change from position to momentum space . The eigenfunctions of
N
{\displaystyle N}
are the usual Hermite functions
ψ
n
(
x
)
{\displaystyle \psi _{n}(x)}
which are therefore also Eigenfunctions of
F
{\displaystyle {\mathcal {F}}}
.[ 7]
Probability version
The result of Mehler can also be linked to probability. For this, the variables should be rescaled as x → x /√2 , y → y /√2 , so as to change from the 'physicist's' Hermite polynomials H (.) (with weight function exp(−x 2 )) to "probabilist's" Hermite polynomials He (.) (with weight function exp(−x 2 /2)). Then, E becomes
1
1
−
ρ
2
exp
(
−
ρ
2
(
x
2
+
y
2
)
−
2
ρ
x
y
2
(
1
−
ρ
2
)
)
=
∑
n
=
0
∞
ρ
n
n
!
H
e
n
(
x
)
H
e
n
(
y
)
.
{\displaystyle {\frac {1}{\sqrt {1-\rho ^{2}}}}\exp \left(-{\frac {\rho ^{2}(x^{2}+y^{2})-2\rho xy}{2(1-\rho ^{2})}}\right)=\sum _{n=0}^{\infty }{\frac {\rho ^{n}}{n!}}~{\mathit {He}}_{n}(x){\mathit {He}}_{n}(y)~.}
The left-hand side here is p (x ,y )/p (x )p (y ) where p (x ,y ) is the bivariate Gaussian probability density function for variables x ,y having zero means and unit variances:
p
(
x
,
y
)
=
1
2
π
1
−
ρ
2
exp
(
−
(
x
2
+
y
2
)
−
2
ρ
x
y
2
(
1
−
ρ
2
)
)
,
{\displaystyle p(x,y)={\frac {1}{2\pi {\sqrt {1-\rho ^{2}}}}}\exp \left(-{\frac {(x^{2}+y^{2})-2\rho xy}{2(1-\rho ^{2})}}\right)~,}
and p (x ), p( y) are the corresponding probability densities of x and y (both standard normal).
There follows the usually quoted form of the result (Kibble 1945)[ 8]
p
(
x
,
y
)
=
p
(
x
)
p
(
y
)
∑
n
=
0
∞
ρ
n
n
!
H
e
n
(
x
)
H
e
n
(
y
)
.
{\displaystyle p(x,y)=p(x)p(y)\sum _{n=0}^{\infty }{\frac {\rho ^{n}}{n!}}~{\mathit {He}}_{n}(x){\mathit {He}}_{n}(y)~.}
This expansion is most easily derived by using the two-dimensional Fourier transform of p (x ,y ) , which is
c
(
i
u
1
,
i
u
2
)
=
exp
(
−
(
u
1
2
+
u
2
2
−
2
ρ
u
1
u
2
)
/
2
)
.
{\displaystyle c(iu_{1},iu_{2})=\exp(-(u_{1}^{2}+u_{2}^{2}-2\rho u_{1}u_{2})/2)~.}
This may be expanded as
exp
(
−
(
u
1
2
+
u
2
2
)
/
2
)
∑
n
=
0
∞
ρ
n
n
!
(
u
1
u
2
)
n
.
{\displaystyle \exp(-(u_{1}^{2}+u_{2}^{2})/2)\sum _{n=0}^{\infty }{\frac {\rho ^{n}}{n!}}(u_{1}u_{2})^{n}~.}
The Inverse Fourier transform then immediately yields the above expansion formula.
This result can be extended to the multidimensional case.[ 8] [ 9] [ 10]
Since Hermite functions ψn are orthonormal eigenfunctions of the Fourier transform ,
F
[
ψ
n
]
(
y
)
=
(
−
i
)
n
ψ
n
(
y
)
,
{\displaystyle {\mathcal {F}}[\psi _{n}](y)=(-i)^{n}\psi _{n}(y)~,}
in harmonic analysis and signal processing , they diagonalize the Fourier operator,
F
[
f
]
(
y
)
=
∫
d
x
f
(
x
)
∑
n
≥
0
(
−
i
)
n
ψ
n
(
x
)
ψ
n
(
y
)
.
{\displaystyle {\mathcal {F}}[f](y)=\int dxf(x)\sum _{n\geq 0}(-i)^{n}\psi _{n}(x)\psi _{n}(y)~.}
Thus, the continuous generalization for real angle α can be readily defined (Wiener , 1929;[ 11] Condon , 1937[ 12] ), the fractional Fourier transform (FrFT), with kernel
F
α
=
∑
n
≥
0
(
−
i
)
2
α
n
/
π
ψ
n
(
x
)
ψ
n
(
y
)
.
{\displaystyle {\mathcal {F}}_{\alpha }=\sum _{n\geq 0}(-i)^{2\alpha n/\pi }\psi _{n}(x)\psi _{n}(y)~.}
This is a continuous family of linear transforms generalizing the Fourier transform , such that, for α = π /2 , it reduces to the standard Fourier transform, and for α = −π /2 to the inverse Fourier transform.
The Mehler formula, for ρ = exp(−iα ), thus directly provides
F
α
[
f
]
(
y
)
=
1
−
i
cot
(
α
)
2
π
e
i
cot
(
α
)
2
y
2
∫
−
∞
∞
e
−
i
(
csc
(
α
)
y
x
−
cot
(
α
)
2
x
2
)
f
(
x
)
d
x
.
{\displaystyle {\mathcal {F}}_{\alpha }[f](y)={\sqrt {\frac {1-i\cot(\alpha )}{2\pi }}}~e^{i{\frac {\cot(\alpha )}{2}}y^{2}}\int _{-\infty }^{\infty }e^{-i\left(\csc(\alpha )~yx-{\frac {\cot(\alpha )}{2}}x^{2}\right)}f(x)\,\mathrm {d} x~.}
The square root is defined such that the argument of the result lies in the interval [−π /2, π /2].
If α is an integer multiple of π , then the above cotangent and cosecant functions diverge. In the limit , the kernel goes to a Dirac delta function in the integrand, δ(x−y) or δ(x+y) , for α an even or odd multiple of π , respectively. Since
F
2
{\displaystyle {\mathcal {F}}^{2}}
[f ] = f (−x ),
F
α
{\displaystyle {\mathcal {F}}_{\alpha }}
[f ] must be simply f (x ) or f (−x ) for α an even or odd multiple of π , respectively.
See also
References
^ Mehler, F. G. (1866), "Ueber die Entwicklung einer Function von beliebig vielen Variabeln nach Laplaceschen Functionen höherer Ordnung" , Journal für die Reine und Angewandte Mathematik (in German) (66): 161– 176, ISSN 0075-4102 , ERAM 066.1720cj (cf. p 174, eqn (18) & p 173, eqn (13) )
^ Erdélyi, Arthur ; Magnus, Wilhelm ; Oberhettinger, Fritz; Tricomi, Francesco G. (1955), Higher transcendental functions. Vol. II , McGraw-Hill (scan : p.194 10.13 (22) )
^ Pauli, W. , Wave Mechanics: Volume 5 of Pauli Lectures on Physics (Dover Books on Physics, 2000) ISBN 0486414620 ; See section 44.
^ The quadratic form in its exponent, up to a factor of −1/2, involves the simplest (unimodular, symmetric) symplectic matrix in Sp(2,R ). That is,
(
x
,
y
)
M
(
x
y
)
,
{\displaystyle (x,y){\mathbf {M} }{\begin{pmatrix}x\\y\end{pmatrix}}~,~}
where
M
≡
csch
(
2
t
)
(
cosh
(
2
t
)
−
1
−
1
cosh
(
2
t
)
)
,
{\displaystyle {\mathbf {M} }\equiv \operatorname {csch} (2t){\begin{pmatrix}\cosh(2t)&-1\\-1&\cosh(2t)\end{pmatrix}}~,}
so it preserves the symplectic metric,
M
T
(
0
1
−
1
0
)
M
=
(
0
1
−
1
0
)
.
{\displaystyle {\mathbf {M} }^{\text{T}}~{\begin{pmatrix}0&1\\-1&0\end{pmatrix}}~{\mathbf {M} }={\begin{pmatrix}0&1\\-1&0\end{pmatrix}}~.}
^ Horvathy, Peter (1979). "Extended Feynman Formula for Harmonic Oscillator". International Journal of Theoretical Physics . 18 (4): 245-250. Bibcode :1979IJTP...18..245H . doi :10.1007/BF00671761 . S2CID 117363885 .
^ Wolf, Kurt B. (1979), Integral Transforms in Science and Engineering , Springer ([1] and [2] ); see section 7.5.10.
^ Celeghini, Enrico; Gadella, Manuel; del Olmo, Mariano A. (2021). "Hermite Functions and Fourier Series" . Symmetry . 13 (5): 853. arXiv :2007.10406 . Bibcode :2021Symm...13..853C . doi :10.3390/sym13050853 .
^ a b Kibble, W. F. (1945). "An extension of a theorem of Mehler's on Hermite polynomials". Mathematical Proceedings of the Cambridge Philosophical Society . 41 (1): 12– 15. Bibcode :1945PCPS...41...12K . doi :10.1017/S0305004100022313 . MR 0012728 . S2CID 121931906 .
^ Slepian, David (1972), "On the symmetrized Kronecker power of a matrix and extensions of Mehler's formula for Hermite polynomials", SIAM Journal on Mathematical Analysis , 3 (4): 606– 616, doi :10.1137/0503060 , ISSN 0036-1410 , MR 0315173
^ Hörmander, Lars (1995). "Symplectic classification of quadratic forms, and general Mehler formulas". Mathematische Zeitschrift . 219 : 413– 449. doi :10.1007/BF02572374 . S2CID 122233884 .
^ Wiener , N (1929), "Hermitian Polynomials and Fourier Analysis", Journal of Mathematics and Physics 8 : 70–73.
^ Condon, E. U. (1937). "Immersion of the Fourier transform in a continuous group of functional transformations", Proc. Natl. Acad. Sci. USA 23 , 158–164. online