This paper was converted on www.awesomepapers.org from LaTeX by an anonymous user.
Want to know more? Visit the Converter page.

Coherent States and Generalized Hermite Polynomials for Fractional statistics - interpolating from fermions to bosons

Satish Ramakrishna [email protected]
Abstract

This article develops the geometric structure that results from the θ\theta-commutator αβeiθβα=1\alpha\beta-e^{i\theta}\beta\alpha=1 that provides a continuous interpolation between the Clifford and Heisenberg algebras. We first demonstrate the most general geometrical picture, applicable to all values of NN. After listing the properties of this Hilbert space, we study the calculus of generalized coherent states that result when ξN=0\xi^{N}=0, for N2N\geq 2, including a calculation of the free-energy for particles of intermediate statistics. Lastly, we solve the generalized harmonic oscillator problem and derive generalized versions of the Hermite polynomials for general NN.

Some remarks are made to connect this study to the case of anyons. This study represents the first steps towards developing an anyonic field theory.

I Introduction & Motivation

In an earlier paperSatish1 , we analyzed the Hilbert space derived from the “commutator”

αβeiθβα=1\displaystyle\alpha\beta-e^{i\theta}\beta\alpha=1 (1)

in the case where θ=2πMN\theta=\frac{2\pi M}{N} where M,NM,N are co-prime non-zero natural numbers and where we also define z=eiθz=e^{i\theta}. This algebra was inspired by the properties of anyons and was intended to interpolate between fermionic and bosonic statistics. In particular, we had considered the general case where the vacuum state had a non-zero eigenvalue. When the vacuum state has a zero eigenvalue, however, we are naturally led to the study of variables ξ\xi, such that ξN=0\xi^{N}=0. Such variables, which may be referred to as generalized Grassmann variables have been the subject of much study in the past Biedenharn ; MacFarlane ; Chaichian , of which the most complete and relevant is Chaichian . The difference is that since our focus is on the full range of fractional statistics between fermions and bosons in 2+1 dimensions, we study a full calculus starting with integration rules to a physically reasonable construction of the path integral for the free energy.

In addition to the generalization of the Grassmann variables mentioned above, there is also a rather simple geometrical picture that emerges from the arithmetic for the operators in the algebra Hallnas . Akin to the fuzzy solid representations used for the angular momentum algebra, we prove that the relevant geometrical picture here is a pancake that goes from a sphere (for N=2N=2) to a plane (for NN\rightarrow\infty).

II Summary of Properties

We are going to, in this paper, analyze several properties of the Hilbert space, in the special and physically interesting case where the vacuum state has zero eigenvalue for the operators βα\beta\alpha, as well as αα\alpha^{\dagger}\alpha. Hence, in the notation of Satish1 , we set λ0=0\lambda_{0}=0.

For general integer NN, we deduce the following properties.

  1. 1.

    States are labeled by their eigenvalues under βα\beta\alpha, i.e.,

    βα|λm=λm|λm\displaystyle\beta\alpha\ket{\lambda_{m}}=\lambda_{m}\ket{\lambda_{m}}
    Eigenvalues:λ0λN=0,λ1=1,λ2=1+z,λ3=1+z+z2,,\displaystyle Eigenvalues:\>\lambda_{0}\equiv\lambda_{N}=0,\lambda_{1}=1,\lambda_{2}=1+z,\lambda_{3}=1+z+z^{2},\>...\>,
    λN1=1+z+z2++zN2,andλN=1+z+z2++zN1=0\displaystyle\lambda_{N-1}=1+z+z^{2}+...+z^{N-2},\>and\>\>\lambda_{N}=1+z+z^{2}+...+z^{N-1}=0
    Eigenstates:|λ0|0,|λ1,|λ2,|λN1,\displaystyle Eigenstates:\>\ket{\lambda_{0}}\equiv\ket{0},\ket{\lambda_{1}},\ket{\lambda_{2}},...\ket{\lambda_{N-1}},\>\>\>\>\>\>\>\>
    |λN|λ0|0\displaystyle\ket{\lambda_{N}}\equiv\ket{\lambda_{0}}\equiv\ket{0}\>\>\>\>\>\>\>\>
    λm|λn=δnm\displaystyle\innerproduct{\lambda_{m}}{\lambda_{n}}=\delta_{nm}\>\>\>\>\>\>\>\>\>\>\>\> (2)

    The eigenvectors can be constructed to be orthogonal, since these are also the eigenvectors of the usual number operator αα\alpha^{\dagger}\alpha.

    We propose to call these states “overons”, since they represent one of the two ways to flip anyons (“over” and “under”). The complex conjugate states would be then called “underons”. In Appendix 1, we study possible dynamical system analogs that might result from such excitations.

  2. 2.

    The actions of the operators are

    α|0α|λNα|λ0=0\displaystyle\alpha\ket{0}\equiv\alpha\ket{\lambda_{N}}\equiv\alpha\ket{\lambda_{0}}=0
    α|λm=λm|λm1m>0\displaystyle\alpha\ket{\lambda_{m}}=\sqrt{\lambda_{m}}\ket{\lambda_{m-1}}\>\>\>\>\>\>m>0
    β|λm1=λm|λm\displaystyle\beta\ket{\lambda_{m-1}}=\sqrt{\lambda_{m}}\ket{\lambda_{m}} (3)
  3. 3.

    A consistent identification is β=αT,α=βT\beta=\alpha^{T},\>\alpha=\beta^{T}, i.e., the commutator is ααTzαTα=1\alpha\alpha^{T}-z\alpha^{T}\alpha=1..

  4. 4.

    When we take the complex conjugate of the basic commutator, we get, by an entirely similar procedure to the above, that the eigenstates of aaa^{\dagger}a^{*} are |λm\ket{\lambda_{m}^{*}}. The chain of reasoning is

    ααTzαTα=1\displaystyle\alpha\alpha^{T}-z\alpha^{T}\alpha=1
    ααz1αα=1\displaystyle\alpha^{*}\alpha^{\dagger}-z^{-1}\alpha^{\dagger}\alpha^{*}=1
    (αTα)|λm=λm|λm\displaystyle(\alpha^{T}\alpha)\ket{\lambda_{m}}=\lambda_{m}\ket{\lambda_{m}}
    αα|λm=λm|λm\displaystyle\alpha^{\dagger}\alpha^{*}\ket{\lambda^{*}_{m}}=\lambda_{m}^{*}\ket{\lambda_{m}^{*}} (4)

    Taking the complex conjugate of the third equation above, we get (|λm)=|λm\left(\ket{\lambda_{m}}\right)^{*}=\ket{\lambda_{m}^{*}}. This leads to the equations

    α|λm=λm|λm1\displaystyle\alpha\ket{\lambda_{m}}=\sqrt{\lambda_{m}}\ket{\lambda_{m-1}}
    α|λm=λm|λm1\displaystyle\alpha^{*}\ket{\lambda_{m}^{*}}=\sqrt{\lambda_{m}^{*}}\ket{\lambda_{m-1}^{*}} (5)

    Continuing, we can now write |λm\ket{\lambda_{m}^{*}} as a linear combination of the |λm\ket{\lambda_{m}}. In fact, it is easy to see, from the geometry of the eigenvectors on the complex plane in Fig. 3, that

    |λm=z(m1)|λm\displaystyle\ket{\lambda_{m}^{*}}=z^{-(m-1)}\ket{\lambda_{m}} (6)

    Using this,

    α|λm=z(m1)λm|λm=λm|λm\displaystyle\alpha\ket{\lambda_{m}^{*}}=z^{-(m-1)}\sqrt{\lambda_{m}}\ket{\lambda_{m}}=\sqrt{\lambda_{m}}\ket{\lambda_{m}^{*}}\>\>\>\>\>\>\>\>\>\>\>\>\>\>
    α|λm=(αT)|λm=(αT|λm)=(z(m1)αT|λm)=zm1λm+1|λm+1\displaystyle\alpha^{\dagger}\ket{\lambda_{m}}=(\alpha^{T})^{*}\ket{\lambda_{m}}=\left(\alpha^{T}\ket{\lambda_{m}^{*}}\right)^{*}=\left(z^{-(m-1)}\alpha^{T}\ket{\lambda_{m}}\right)^{*}=z^{m-1}\sqrt{\lambda_{m+1}^{*}}\ket{\lambda_{m+1}^{*}}
    =z1λm+1|λm+1\displaystyle=z^{-1}\sqrt{\lambda_{m+1}^{*}}\ket{\lambda_{m+1}}\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\> (7)

    Using this, and defining λm|\bra{\lambda_{m}} as the usual hermitian conjugate transpose of |λm\ket{\lambda_{m}},

    αα|λm=|λm+1||λm\displaystyle\alpha\alpha^{\dagger}\ket{\lambda_{m}}=|\lambda_{m+1}|\ket{\lambda_{m}}\>\>\>\>\>\>\>
    αα|λm=|λm||λm\displaystyle\alpha^{\dagger}\alpha\ket{\lambda_{m}}=|\lambda_{m}|\ket{\lambda_{m}}
    λm|αααα|λm=|λm+1||λm|\displaystyle\rightarrow\>\>\bra{\lambda_{m}}\alpha\alpha^{\dagger}-\alpha^{\dagger}\alpha\ket{\lambda_{m}}=|\lambda_{m+1}|-|\lambda_{m}| (8)

    The traditional “number” operator αα\alpha^{\dagger}\alpha is diagonal in the same basis that αTα\alpha^{T}\alpha is. Since αα\alpha^{\dagger}\alpha is a hermitian operator, it is consistent that the |λm\ket{\lambda_{m}} is an orthonormal basis Banks .

  5. 5.

    The eigenvalue spectrum of βααTα\beta\alpha\equiv\alpha^{T}\alpha (magnitude as well as complex vectors) is as below and in reference Satish1 and is plotted in Fig. 1. Note that we have θ=2πMN\theta=\frac{2\pi M}{N} and we have used M=1M=1 for the graphs in Fig. 1.

    λ0=0\displaystyle\lambda_{0}=0\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>
    λm=z0+z1++zm1=1eimθ1eiθ=ei(m1)θ2sin(mθ2)sin(θ2)\displaystyle\lambda_{m}=z^{0}+z^{1}+...+z^{m-1}=\frac{1-e^{im\theta}}{1-e^{i\theta}}=e^{i\frac{(m-1)\theta}{2}}\frac{\sin{\frac{m\theta}{2}}}{\sin{\frac{\theta}{2}}} (9)
    Figure 1: Eigenvalue Spectrum
    Refer to caption
    Refer to caption
  6. 6.

    The matrices α\alpha and β=αT\beta=\alpha^{T} are displayed explicitly below, for θ=2πMN\theta=\frac{2\pi M}{N}. They are N×NN\times N matrices, since the eigenstates are NN-dimensional vectors. The eigenvectors are as below

    |λ0=(𝟏000...00),|λ1=(0𝟏00...00),|λ2=(00𝟏0...00)|λN1=(0000...0𝟏)\displaystyle\ket{\lambda_{0}}=\left(\begin{array}[]{c}{\bf 1}\\ 0\\ 0\\ 0\\ .\\ .\\ .\\ 0\\ 0\end{array}\right)\>\>\>,\>\>\ket{\lambda_{1}}=\left(\begin{array}[]{c}0\\ {\bf 1}\\ 0\\ 0\\ .\\ .\\ .\\ 0\\ 0\end{array}\right)\>\>\>,\>\>\ket{\lambda_{2}}=\left(\begin{array}[]{c}0\\ 0\\ {\bf 1}\\ 0\\ .\\ .\\ .\\ 0\\ 0\end{array}\right)\>\>\>...\>\>\ket{\lambda_{N-1}}=\left(\begin{array}[]{c}0\\ 0\\ 0\\ 0\\ .\\ .\\ .\\ 0\\ {\bf 1}\end{array}\right)\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\> (46)

    while the operators are

    α=(𝟎   1   0   0   0   000   01+z   0   0   000   0   01+z+z2   0   000   0   0   01+z+z2+z3   00...0   0   0   0   0   01+z+z2++zN10   0   0   0   0   0𝟎)\displaystyle\alpha=\left(\begin{array}[]{cccccccc}{\bf 0}&\>\>\>1&\>\>\>0&\>\>\>0&\>\>\>0&...&\>\>\>0&0\\ 0&\>\>\>{\bf 0}&\>\>\>\sqrt{1+z}&\>\>\>0&\>\>\>0&...&\>\>\>0&0\\ 0&\>\>\>0&\>\>\>{\bf 0}&\>\>\>\sqrt{1+z+z^{2}}&\>\>\>0&...&\>\>\>0&0\\ 0&\>\>\>0&\>\>\>0&\>\>\>{\bf 0}&\>\>\>\sqrt{1+z+z^{2}+z^{3}}&...&\>\>\>0&0\\ .\\ .\\ .\\ 0&\>\>\>0&\>\>\>0&\>\>\>0&\>\>\>0&...&\>\>\>{\bf 0}&\sqrt{1+z+z^{2}+...+z^{N-1}}\\ 0&\>\>\>0&\>\>\>0&\>\>\>0&\>\>\>0&...&\>\>\>0&{\bf 0}\end{array}\right) (56)
    β=αT=(𝟎   0   0   0   0   001   0   0   0   0   0001+z   0   0   0   000   01+z+z2   0   0   000   0   01+z+z2+z3   0   00...0   0   0   0   01+z+z2++zN1𝟎)\displaystyle\beta=\alpha^{T}=\left(\begin{array}[]{cccccccc}{\bf 0}&\>\>\>0&\>\>\>0&\>\>\>0&\>\>\>0&...&\>\>\>0&0\\ 1&\>\>\>{\bf 0}&\>\>\>0&\>\>\>0&\>\>\>0&...&\>\>\>0&0\\ 0&\>\>\>\sqrt{1+z}&\>\>\>{\bf 0}&\>\>\>0&\>\>\>0&...&\>\>\>0&0\\ 0&\>\>\>0&\>\>\>\sqrt{1+z+z^{2}}&\>\>\>{\bf 0}&\>\>\>0&...&\>\>\>0&0\\ 0&\>\>\>0&\>\>\>0&\>\>\>\sqrt{1+z+z^{2}+z^{3}}&\>\>\>{\bf 0}&...&\>\>\>0&0\\ .\\ .\\ .\\ 0&\>\>\>0&\>\>\>0&\>\>\>0&\>\>\>0&...&\sqrt{1+z+z^{2}+...+z^{N-1}}&{\bf 0}\end{array}\right) (66)

    while the commutator is

    ααTαTα=αββα=(𝟏   0   0   0   000𝐳   0   0   0   000   0𝐳𝟐   0   0   000   0   0𝐳𝟑   0   000   0   0   0𝐳𝟒   00...0   0   0   0   0𝐳𝐍𝟐   00   0   0   0   0   0𝐳𝐍𝟏)\displaystyle\alpha\alpha^{T}-\alpha^{T}\alpha=\alpha\beta-\beta\alpha=\left(\begin{array}[]{cccccccc}{\bf 1}&\>\>\>0&\>\>\>0&\>\>\>0&\>\>\>0&...&0\\ 0&\>\>\>{\bf z}&\>\>\>0&\>\>\>0&\>\>\>0&...&\>\>\>0&0\\ 0&\>\>\>0&\>\>\>{\bf z^{2}}&\>\>\>0&\>\>\>0&...&\>\>\>0&0\\ 0&\>\>\>0&\>\>\>0&\>\>\>{\bf z^{3}}&\>\>\>0&...&\>\>\>0&0\\ 0&\>\>\>0&\>\>\>0&\>\>\>0&\>\>\>{\bf z^{4}}&...&\>\>\>0&0\\ .\\ .\\ .\\ 0&\>\>\>0&\>\>\>0&\>\>\>0&\>\>\>0&...&{\bf z^{N-2}}&\>\>\>0\\ 0&\>\>\>0&\>\>\>0&\>\>\>0&\>\>\>0&...&\>\>\>0&{\bf z^{N-1}}\end{array}\right)\>\>\>\>\>\>\>\>\>\>\>\> (77)
    =diag[1,z,z2,z3,,zN1]\displaystyle=diag\left[1,z,z^{2},z^{3},...,z^{N-1}\right]\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\> (78)

    and

    αααα=diag[10,|λ2||λ1|,|λ3||λ2|,,|λN||λN1|]\displaystyle\alpha\alpha^{\dagger}-\alpha^{\dagger}\alpha=diag\left[1-0,|\lambda_{2}|-|\lambda_{1}|,|\lambda_{3}|-|\lambda_{2}|,...,|\lambda_{N}|-|\lambda_{N-1}|\right]
    =diag[1,sin2θ2sinθ2sinθ2,sin3θ2sin2θ2sinθ2,sin4θ2sin3θ2sinθ2,,sin(N1)θ2sinθ2]\displaystyle=diag[1,\frac{\sin\frac{2\theta}{2}-\sin\frac{\theta}{2}}{\sin\frac{\theta}{2}},\frac{\sin\frac{3\theta}{2}-\sin\frac{2\theta}{2}}{\sin\frac{\theta}{2}},\frac{\sin\frac{4\theta}{2}-\sin\frac{3\theta}{2}}{\sin\frac{\theta}{2}},...,\frac{-\sin\frac{(N-1)\theta}{2}}{\sin\frac{\theta}{2}}]\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\> (79)

    and

    αα𝒞αα=1\displaystyle\alpha\alpha^{\dagger}-{\mathcal{C}}\alpha^{\dagger}\alpha=1
    𝒞=diag[|λ1|1|λ0|,|λ2|1|λ1|,|λ3|1|λ2|,,|λN|1|λN1|]\displaystyle{\mathcal{C}}=diag[\frac{|\lambda_{1}|-1}{|\lambda_{0}|},\frac{|\lambda_{2}|-1}{|\lambda_{1}|},\frac{|\lambda_{3}|-1}{|\lambda_{2}|},...,\frac{|\lambda_{N}|-1}{|\lambda_{N-1}|}] (80)

    The top-left component |λ1|1|λ0|\frac{|\lambda_{1}|-1}{|\lambda_{0}|} of 𝒞\mathcal{C} is not determined by the above commutator since αα=0\alpha^{\dagger}\alpha=0 for the state |λ0\ket{\lambda_{0}}. However, we can determine it by taking the limit of the expressions for λ00\lambda_{0}\rightarrow 0 as in Satish1 ; it becomes cosθ\cos\theta, which is 1-1 for the fermion limit and +1+1 for the bosonic limit. Hence 𝒞\mathcal{C} is -\cal I (the identity matrix) for fermions and ++\cal I for bosons.

  7. 7.

    When we compute scattering amplitude matrix elements for different particles, we will have to re-order the annihilation and creation operators, then will be left with a product of terms like αmβm\alpha^{m}\beta^{m}, i.e.,

    αβ=(1+zβα)\displaystyle\alpha\beta=(1+z\>\beta\alpha)
    α2β2=(1+z)+(z+2z+z3)βα+z4β2α2\displaystyle\alpha^{2}\beta^{2}=(1+z)+(z+2z+z^{3})\beta\alpha+z^{4}\beta^{2}\alpha^{2}\>
    \displaystyle...

    When we compute expectation values in the vacuum state (i.e., 0|αmβm|0\bra{0}\alpha^{m}\beta^{m}\ket{0}), only the constant terms will be left and they are, for the first few powers

    m=1   1\displaystyle m=1\>\>\>\rightarrow\>\>\>1
    m=2   1+z\displaystyle m=2\>\>\>\rightarrow\>\>\>1+z
    m=3   1+2z+2z2+z3\displaystyle m=3\>\>\>\rightarrow\>\>\>1+2z+2z^{2}+z^{3}
    m=4   1+3z+5z2+6z3+5z4+3z5+z6\displaystyle m=4\>\>\>\rightarrow\>\>\>1+3z+5z^{2}+6z^{3}+5z^{4}+3z^{5}+z^{6}
    m=5   1+4z+9z2+15z3+20z4+22z5+20z6+15z7+9z8+4z9+z10\displaystyle m=5\>\>\>\rightarrow\>\>\>1+4z+9z^{2}+15z^{3}+20z^{4}+22z^{5}+20z^{6}+15z^{7}+9z^{8}+4z^{9}+z^{10}
    m=6(1,5,14,29,49,71,90,101,101,90,71,49,29,14,5,1)\displaystyle m=6\>\>\>\rightarrow(1,5,14,29,49,71,90,101,101,90,71,49,29,14,5,1)\>\>\>\>\>\>\>\>\>\>\>\>\>\>\> (81)

    where we have represented the polynomial by just its coefficients (the Mahonian numbers Mahon ) in the last case. As can be checked quickly, these are the polynomials λ1,λ2×λ1,λ3×λ2×λ1\lambda_{1},\lambda_{2}\times\lambda_{1},\lambda_{3}\times\lambda_{2}\times\lambda_{1} etc.

III Geometrical Interpretation

The most general geometrical construction is

X=α+α2,Y=αα2i,Z=12[α,α]\displaystyle X=\frac{\alpha+\alpha^{\dagger}}{2}\;,\;Y=\frac{\alpha-\alpha^{\dagger}}{2i}\>,\>Z=\frac{1}{2}[\alpha,\alpha^{\dagger}]
X2+Y2=12(αα+αα)\displaystyle X^{2}+Y^{2}=\frac{1}{2}\left(\alpha\alpha^{\dagger}+\alpha^{\dagger}\alpha\right)
2Z=(αααα)\displaystyle 2Z=\left(\alpha\alpha^{\dagger}-\alpha^{\dagger}\alpha\right)
m,n(αα+αα)m,n=(|λm+1|+|λm|)δm,n\displaystyle{\cal L}_{m,n}\equiv\left(\alpha\alpha^{\dagger}+\alpha^{\dagger}\alpha\right)_{m,n}=\left(|\lambda_{m+1}|+|\lambda_{m}|\right)\delta_{m,n}
m,n(αααα)m,n=(|λm+1||λm|)δm,n\displaystyle{\cal M}_{m,n}\equiv\left(\alpha\alpha^{\dagger}-\alpha^{\dagger}\alpha\right)_{m,n}=\left(|\lambda_{m+1}|-|\lambda_{m}|\right)\delta_{m,n} (82)

and observe that

((|λm+1|+|λm|)sinθ4)2+((|λm+1||λm|)cosθ4)2=1\displaystyle\bigg{(}(|\lambda_{m+1}|+|\lambda_{m}|)\sin\frac{\theta}{4}\bigg{)}^{2}+\bigg{(}(|\lambda_{m+1}|-|\lambda_{m}|)\cos\frac{\theta}{4}\bigg{)}^{2}=1 (83)

we obtain the equation

(X2+Y2)2sin2θ4+Z2cos2θ4=14\displaystyle(X^{2}+Y^{2})^{2}\sin^{2}\frac{\theta}{4}+Z^{2}\cos^{2}\frac{\theta}{4}=\frac{1}{4} (84)

This equation can be re-phrased as an invariant of the group that underlies the algebra. The algebra can be written most simply with the creation/annihilation operators as

[α,α]=2Z\displaystyle\left[\alpha,\alpha^{\dagger}\right]=2Z
[α,Z]=SDα\displaystyle\left[\alpha,Z\right]=S_{D}\>\alpha
[α,Z]=αSD\displaystyle\left[\alpha^{\dagger},Z\right]=-\alpha^{\dagger}\>S_{D}
tan2θ4=1221\displaystyle\tan^{2}\frac{\theta}{4}=\frac{1-{\cal M}^{2}}{{\cal L}^{2}-1} (85)

In Equation (17), the two limits θ0\theta\rightarrow 0 and θπ\theta\rightarrow\pi are consistent on both sides of the equation. Additionally, {\cal L} and {\cal M} are the matrices as defined in Equation (23) and SDS_{D} is the real, diagonal matrix

SD=diag[|λ0|2+|λ2|2|λ1|,|λ1|2+|λ3|2|λ2|,,|λN2|2+|λN|2|λN1|]\displaystyle S_{D}=diag\left[\frac{|\lambda_{0}|}{2}+\frac{|\lambda_{2}|}{2}-|\lambda_{1}|,\frac{|\lambda_{1}|}{2}+\frac{|\lambda_{3}|}{2}-|\lambda_{2}|,...,\frac{|\lambda_{N-2}|}{2}+\frac{|\lambda_{N}|}{2}-|\lambda_{N-1}|\right]\
Dm2(|λ|)\displaystyle\equiv D_{m}^{2}(|\lambda|)\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\> (86)

which is the finite-difference Laplacian of a diagonal matrix with absolute values of the eigenvalues along the diagonal. In this notation, from Equation (17), Z=Dm(|λ|)Z=D_{m}(|\lambda|), so that the commutator in that equation can be written in the interesting form

[α,Dm(|λ|)]=Dm2(|λ|)α\displaystyle\left[\alpha,D_{m}(|\lambda|)\right]=D_{m}^{2}(|\lambda|)\alpha (87)

Incidentally, for N=3N=3, this is consistent with the previous paper’sSatish1 Equation (25) at the points resolved within the fuzzy ellipsoid (Z=Jz=0,±12Z=J_{z}=0,\pm\frac{1}{2}). In this case,

Jz=12(𝟏000𝟎000𝟏),Jx=α+α2=12(0𝟏0𝟏0𝟏+𝐞𝐢𝟐𝐌π𝟑0𝟏+𝐞𝐢𝟐π𝐌𝟑0),\displaystyle J_{z}=\frac{1}{2}\left(\begin{array}[]{ccc}{\bf 1}&0&0\\ 0&{\bf 0}&0\\ 0&0&{\bf-1}\end{array}\right)\>,\>J_{x}=\frac{\alpha+\alpha^{\dagger}}{2}=\frac{1}{2}\left(\begin{array}[]{ccc}0&{\bf 1}&0\\ {\bf 1}&0&{\bf\sqrt{1+e^{i\frac{2M\pi}{3}}}}\\ 0&{\bf\sqrt{1+e^{-i\frac{2\pi M}{3}}}}&0\end{array}\right)\>, (94)
Jy=αα2i=i2(0𝟏0𝟏0𝟏+𝐞𝐢𝟐𝐌π𝟑0𝟏+𝐞𝐢𝟐π𝐌𝟑0)\displaystyle\>J_{y}=\frac{\alpha-\alpha^{\dagger}}{2i}=\frac{i}{2}\left(\begin{array}[]{ccc}0&{\bf-1}&0\\ {\bf 1}&0&-\bf{\sqrt{1+e^{i\frac{2M\pi}{3}}}}\\ 0&\bf{\sqrt{1+e^{-i\frac{2\pi M}{3}}}}&0\end{array}\right) (98)
Jz2=14(𝟏000𝟎000𝟏),Jx2+Jy2=14(𝟐000𝟒000𝟐)\displaystyle J_{z}^{2}=\frac{1}{4}\left(\begin{array}[]{ccc}{\bf 1}&0&0\\ 0&{\bf 0}&0\\ 0&0&{\bf 1}\end{array}\right)\>,\>J_{x}^{2}+J_{y}^{2}=\frac{1}{4}\left(\begin{array}[]{ccc}{\bf 2}&0&0\\ 0&{\bf 4}&0\\ 0&0&{\bf 2}\end{array}\right)\>\>\>\>\>\>\>\>\>\>\>\>\> (105)

satisfies both the Equations (25) and following in the previous paper Satish1 , i.e.,

Jx2+Jy2+2Jz2=1&(Jx2+Jy2)2+3Jz2=1\displaystyle J_{x}^{2}+J_{y}^{2}+2J_{z}^{2}=1\>\>\>\>\&\>\>\>\>(J_{x}^{2}+J_{y}^{2})^{2}+3J_{z}^{2}=1 (106)

We have already noted the equivalence for N=2N=2.

The surface described by the above equation is pancake-shaped aligned along the z-axes, as in Fig. 2.

Figure 2: Geometrical Interpretation of the eigenvalue surface
Refer to caption

For large NN, we can write the solutions to the above equation (in terms of the polar radius coordinate ρ\rho) as

ρ2=X2+Y2=(m+12)mq212(1+3m+2m2)\displaystyle\rho^{2}=X^{2}+Y^{2}=(m+\frac{1}{2})-\frac{mq^{2}}{12}(1+3m+2m^{2})
z=12mq24(1+m)\displaystyle z=\frac{1}{2}-\frac{mq^{2}}{4}(1+m)\>\>\>\>\>\>\>\>\>\>\>\>\>\>\> (107)

which yields concentric circular strips on the plane z=12z=\frac{1}{2}, where each state has radius m+12\propto\sqrt{m+\frac{1}{2}}. This is the usual picture for Landau levels; here this is a geometrical representation of the usual bosonic levels.

To show how a spherical object for N=2N=2 transforms into the flat plane in the NN\rightarrow\infty limit, we can compute the surface area of the pancake. The area integral is computed (define a=sinθ4a=\sin\frac{\theta}{4}) as

𝒜(a)=2(2π)ρ=0ρ=12a𝑑ρρ1+(dzdρ)2\displaystyle{\cal A}(a)=2(2\pi)\int_{\rho=0}^{\rho=\frac{1}{\sqrt{2a}}}\>d\rho\>\rho\sqrt{1+(\frac{dz}{d\rho})^{2}}
=2πal=0l=12𝑑l1+4al3(1a2)(14l2)\displaystyle=\frac{2\pi}{a}\int_{l=0}^{l=\frac{1}{2}}\>dl\sqrt{1+\frac{4al^{3}}{(1-a^{2})(\frac{1}{4}-l^{2})}} (108)

Clearly, this diverges as a0a\rightarrow 0, which corresponds to NN\rightarrow\infty, since θ=2πN\theta=\frac{2\pi}{N}.

We plot the area as a function of aa in Fig. 3. Indeed, the pancake like closed surface turns into the infinite plane as NN\rightarrow\infty.

Figure 3: Area of the geometrical surface
Refer to caption

IV Coherent States for general NN - Generalized Grassmann variables

For a general NN-type algebra, we can write down an eigenstate for α\alpha as

α|ξ=ξ|ξ\displaystyle\alpha\ket{\xi}=\xi\ket{\xi}\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>
|ξ=|λ0+ξλ1|λ1+ξ2λ2λ1|λ2+ξ3λ3λ2λ1|λ3+\displaystyle\ket{\xi}=\ket{\lambda_{0}}+\frac{\xi}{\sqrt{\lambda_{1}}}\ket{\lambda_{1}}+\frac{\xi^{2}}{\sqrt{\lambda_{2}\lambda_{1}}}\ket{\lambda_{2}}+\frac{\xi^{3}}{\sqrt{\lambda_{3}\lambda_{2}\lambda_{1}}}\ket{\lambda_{3}}+\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>
+ξN2λN2λN3λ1|λN2+ξN1λN1λN2λ1|λN1\displaystyle...+\frac{\xi^{N-2}}{\sqrt{\lambda_{N-2}\lambda_{N-3}...\lambda_{1}}}\ket{\lambda_{N-2}}+\frac{\xi^{N-1}}{\sqrt{\lambda_{N-1}\lambda_{N-2}...\lambda_{1}}}\ket{\lambda_{N-1}}\>\>\>\>\>\>\>\>\> (109)

The above expansion is identical to the usual formula for boson coherent states in the limit NN\rightarrow\infty, as well as for fermion coherent states at N=2N=2 (the usual Grassmann variables). We also define the “bra” vector as the adjoint vector, where ξ\xi^{\dagger} is the adjoint of ξ\xi and is independent of ξ\xi, i.e.,

ξ|=λ0|+ξλ1λ1|+(ξ)2λ2λ1λ2|+(ξ)3λ3λ2λ1λ3|+\displaystyle\bra{\xi}=\bra{\lambda_{0}}+\frac{\xi^{\dagger}}{\sqrt{\lambda_{1}^{*}}}\bra{\lambda_{1}}+\frac{(\xi^{\dagger})^{2}}{\sqrt{\lambda_{2}^{*}\lambda_{1}^{*}}}\bra{\lambda_{2}}+\frac{(\xi^{\dagger})^{3}}{\sqrt{\lambda_{3}^{*}\lambda_{2}^{*}\lambda_{1}^{*}}}\bra{\lambda_{3}}+\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>
+(ξ)N2λN2λN3λ1λN2|+(ξ)N1λN1λN2λ1λN1|\displaystyle...+\frac{(\xi^{\dagger})^{N-2}}{\sqrt{\lambda_{N-2}^{*}\lambda_{N-3}^{*}...\lambda_{1}^{*}}}\bra{\lambda_{N-2}}+\frac{(\xi^{\dagger})^{N-1}}{\sqrt{\lambda_{N-1}^{*}\lambda_{N-2}^{*}...\lambda_{1}^{*}}}\bra{\lambda_{N-1}}
ξ|α=ξ|ξ\displaystyle\bra{\xi}\alpha^{\dagger}=\bra{\xi}\xi^{\dagger}\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\> (110)

That they are eigenvectors may be checked by applying α\alpha to the state and using Equation (3), we postulate that ξ\xi and α\alpha commute. Additionally, we posit that ξN=0\xi^{N}=0 and the transposition relations (the matrix 𝒞0{\cal C}_{0} is as defined in Equation (12))

(ξ)=ξ,ξξ=𝒞0ξξ,ξξ=𝒞01ξξ(\xi^{\dagger})^{\dagger}=\xi\>\>,\>\>\xi\xi^{\dagger}={\cal C}_{0}\>\xi^{\dagger}\xi\>\>,\>\>\xi^{\dagger}\xi={\cal C}_{0}^{-1}\>\xi\>\xi^{\dagger} (111)

What sort of object is ξ\xi? The two relations above make sense only if ξ\xi were itself an N×NN\times N matrix, for consistency, we assume is a direct product of a “generalized Grassmann matrix” and a unit N×NN\times N matrix in the states’ eigenbasis, hence, commutes with all other complex number matrices in the same eigenbasis. Hence ξT\xi^{T} and ξ\xi^{\dagger} are all reasonable objects to define. The above statements and equations are also consistent with the statement that a term like ξξ\xi^{\dagger}\xi is “real”. This is true, since 𝒞\cal C is a real matrix. In addition, since ξα\xi\sim\alpha, it is consistent to require the equations below (in line with Equation (4)),

ξξT=zξTξ\displaystyle\xi\xi^{T}=z\xi^{T}\xi
ξξ=𝒞0ξξ\displaystyle\xi\xi^{\dagger}={\cal C}_{0}\xi^{\dagger}\xi

The description of this algebra is similar to the treatment in reference Chaichian , however, the difference here is that these variables are directly coherent state variables, as in the usual definition Murayama1 .

Some auxiliary results are written below. Again, note that all these products are scalars times the unit matrix.

ξ|ξ=1+ξξ|λ1|+(ξ)2ξ2|λ2λ1|+(ξ)3ξ3|λ3λ2λ1|++(ξ)N1ξN1|λN1λN2λ1|\displaystyle\innerproduct{\xi}{\xi}=1+\frac{\xi^{\dagger}\xi}{|\lambda_{1}|}+\frac{(\xi^{\dagger})^{2}\xi^{2}}{|\lambda_{2}\lambda_{1}|}+\frac{(\xi^{\dagger})^{3}\xi^{3}}{|\lambda_{3}\lambda_{2}\lambda_{1}|}+...+\frac{(\xi^{\dagger})^{N-1}\xi^{N-1}}{|\lambda_{N-1}\lambda_{N-2}...\lambda_{1}|} (113)

Note that in the NN\rightarrow\infty limit, when z1z\rightarrow 1, ξ|ξ=eξξ\innerproduct{\xi}{\xi}=e^{\xi^{\dagger}\xi}, which is appropriate for bosonic coherent states. Taking the limit in the opposite direction, the behavior is appropriate for fermions, when N=2N=2 the algebra automatically yields ξ|ξ=eξξ=eξξ\innerproduct{\xi}{\xi}=e^{\xi^{\dagger}\xi}=e^{-\xi\xi^{\dagger}}.

We can now write down, from the expansion in Equation (28), a series expansion for 1ξ|ξ\frac{1}{\innerproduct{\xi}{\xi}}, which also terminates at the term (ξ)N1ξN1(\xi^{\dagger})^{N-1}\xi^{N-1}, since higher powers are 0.

The following integrals are postulated, in order to match the boundary cases for bosonic variables as well as for fermionic (N=2)(N=2) Grassmanns. We set

𝑑ξ𝑑ξ1ξ|ξ=1Normalization\displaystyle\int d\xi^{\dagger}d\xi\frac{1}{\innerproduct{\xi}{\xi}}=1\>\>\>\>\>Normalization\>\>
𝑑ξ𝑑ξ1ξ|ξξξ=𝒞0|λ1|FirstMoment\displaystyle\int d\xi^{\dagger}d\xi\frac{1}{\innerproduct{\xi}{\xi}}\xi^{\dagger}\xi={\cal C}_{0}|\lambda_{1}|\>\>\>\>\>First\>Moment\>\>
𝑑ξ𝑑ξ1ξ|ξ(ξ)2ξ2=𝒞02|λ2λ1|SecondMoment\displaystyle\int d\xi^{\dagger}d\xi\frac{1}{\innerproduct{\xi}{\xi}}(\xi^{\dagger})^{2}\xi^{2}={\cal C}_{0}^{2}|\lambda_{2}\lambda_{1}|\>\>\>\>\>Second\>Moment\>\>
\displaystyle...
𝑑ξ𝑑ξ1ξ|ξ(ξ)nξn=𝒞0n|λnλn1λ1|nthMoment\displaystyle\int d\xi^{\dagger}d\xi\frac{1}{\innerproduct{\xi}{\xi}}(\xi^{\dagger})^{n}\xi^{n}={\cal C}_{0}^{n}|\lambda_{n}\lambda_{n-1}...\lambda_{1}|\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>n^{th}\>Moment\>\>\>
\displaystyle...
𝑑ξ𝑑ξ1ξ|ξ(ξ)N1ξN1=𝒞0N1|λN1λN2λ1|(N1)thMoment\displaystyle\int d\xi^{\dagger}d\xi\frac{1}{\innerproduct{\xi}{\xi}}(\xi^{\dagger})^{N-1}\xi^{N-1}={\cal C}_{0}^{N-1}|\lambda_{N-1}\lambda_{N-2}...\lambda_{1}|\>\>\>\>\>(N-1)^{th}\>Moment\>\>\> (114)

which is consistent with

𝑑ξ𝑑ξ  1=0\displaystyle\int d\xi^{\dagger}d\xi\>\>1=0
𝑑ξ𝑑ξξξ=0\displaystyle\int d\xi^{\dagger}d\xi\>\>\xi^{\dagger}\xi=0
𝑑ξ𝑑ξ(ξ)2ξ2=0\displaystyle\int d\xi^{*}{\dagger}d\xi\>\>(\xi^{\dagger})^{2}\xi^{2}=0
\displaystyle...
𝑑ξ𝑑ξ(ξ)N1ξN1=𝒞0N1|λN1λN2λ1|\displaystyle\int d\xi^{\dagger}d\xi\>\>(\xi^{\dagger})^{N-1}\xi^{N-1}={\cal C}_{0}^{N-1}|\lambda_{N-1}\lambda_{N-2}...\lambda_{1}| (115)

While the first (moment) variety of integral can be checked for the limiting fermion and boson cases, the second variety of integrals (without the normalization) cannot be properly defined in the bosonic cases since it isn’t convergent. The first method can be treated as a regularized integral.

The one-variable version of these integrals can be defined, in consistency with the above equations, as

𝑑ξξ=0\displaystyle\int d\xi\>\xi=0
𝑑ξξ2=0\displaystyle\int d\xi\>\xi^{2}=0
\displaystyle...
𝑑ξξN1=𝒞0N1|λN1λN2λ1|\displaystyle\int d\xi\>\xi^{N-1}={\cal C}_{0}^{N-1}\sqrt{|\lambda_{N-1}\lambda_{N-2}...\lambda_{1}|}
𝑑ξ(ξ)N1=𝒞0N1|λN1λN2λ1|\displaystyle\int d\xi^{\dagger}\>(\xi^{\dagger})^{N-1}={\cal C}_{0}^{N-1}\sqrt{|\lambda_{N-1}\lambda_{N-2}...\lambda_{1}|} (116)

It is possible to define an identity operator, so that (using Equation (29)),

=𝑑ξ𝑑ξ1ξ|ξ|𝒞01ξξ|\displaystyle\mathcal{I}=\int d\xi^{\dagger}d\xi\>\frac{1}{\innerproduct{\xi}{\xi}}\ket{{\cal C}_{0}^{-1}\xi^{\dagger}}\bra{\xi^{\dagger}}\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>
=dξdξ1ξ|ξ(|λ0λ0|+𝒞01ξξ|λ1||λ1λ1|+(𝒞01ξ𝒞01ξ)ξ2|λ2λ1||λ2λ2|+\displaystyle=\int d\xi^{\dagger}d\xi\>\frac{1}{\innerproduct{\xi}{\xi}}\bigg{(}\ket{\lambda_{0}}\bra{\lambda_{0}}+\frac{{\cal C}_{0}^{-1}\xi^{\dagger}\xi}{|\lambda_{1}|}\ket{\lambda_{1}}\bra{\lambda_{1}}+\frac{({\cal C}_{0}^{-1}\xi^{\dagger}{\cal C}_{0}^{-1}\xi^{\dagger})\xi^{2}}{|\lambda_{2}\lambda_{1}|}\ket{\lambda_{2}}\bra{\lambda_{2}}+...
+(𝒞01ξ𝒞01ξ)ξN1|λN1λ1||λN1λN1|)\displaystyle+\frac{({\cal C}_{0}^{-1}\xi^{\dagger}...{\cal C}_{0}^{-1}\xi^{\dagger})\xi^{N-1}}{|\lambda_{N-1}...\lambda_{1}|}\ket{\lambda_{N-1}}\bra{\lambda_{N-1}}\bigg{)}
=|λ0λ0|+|λ1λ1|++|λN1λN1|\displaystyle\rightarrow\mathcal{I}=\ket{\lambda_{0}}\bra{\lambda_{0}}+\ket{\lambda_{1}}\bra{\lambda_{1}}+...+\ket{\lambda_{N-1}}\bra{\lambda_{N-1}}\>\>\>\>\>\>\>\>\>\>\>\>\> (117)

V The Trace and Path Integrals

The trace of an operator AA that commutes with and can be taken through ξ\xi is

Tr(𝐀)=𝑑ξ𝑑ξ1ξ|ξ𝒞01ξ|𝐀|ξ\displaystyle Tr({\bf A})=\int d\xi^{\dagger}d\xi\frac{1}{\innerproduct{\xi}{\xi}}\bra{{\cal C}_{0}^{-1}\xi}{{\bf A}}\ket{\xi}\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>
=dξdξ1ξ|ξ(λ0|+𝒞01ξλ1λ1|+(𝒞01ξ)(𝒞01ξ)λ2λ1λ2|++(𝒞01ξ)(𝒞01ξ)λN2λN3λ1λN2|+\displaystyle=\int d\xi^{\dagger}d\xi\frac{1}{\innerproduct{\xi}{\xi}}\>(\bra{\lambda_{0}}+\frac{{\cal C}_{0}^{-1}\>\xi^{\dagger}}{\sqrt{\lambda_{1}^{*}}}\bra{\lambda_{1}}+\frac{({\cal C}_{0}^{-1}\>\xi^{\dagger})({\cal C}_{0}^{-1}\>\xi^{\dagger})}{\sqrt{\lambda_{2}^{*}\lambda_{1}^{*}}}\bra{\lambda_{2}}+...+\frac{({\cal C}_{0}^{-1}\>\xi^{\dagger})...({\cal C}_{0}^{-1}\>\xi^{\dagger})}{\sqrt{\lambda_{N-2}^{*}\lambda_{N-3}^{*}...\lambda_{1}^{*}}}\bra{\lambda_{N-2}}+
(𝒞01ξ)(𝒞01ξ)λN1λN2λ1λN1|)𝐀(|λ0+ξλ1|λ1+ξ2λ2λ1|λ2+\displaystyle\frac{({\cal C}_{0}^{-1}\>\xi^{\dagger})...({\cal C}_{0}^{-1}\>\xi^{\dagger})}{\sqrt{\lambda_{N-1}^{*}\lambda_{N-2}^{*}...\lambda_{1}^{*}}}\bra{\lambda_{N-1}}){\bf A}(\ket{\lambda_{0}}+\frac{\xi}{\sqrt{\lambda_{1}}}\ket{\lambda_{1}}+\frac{\xi^{2}}{\sqrt{\lambda_{2}\lambda_{1}}}\ket{\lambda_{2}}+\>\>\>\>\>\>
+ξN2λN2λN3λ1|λN2+ξN1λN1λN2λ1|λN1)\displaystyle...+\frac{\xi^{N-2}}{\sqrt{\lambda_{N-2}\lambda_{N-3}...\lambda_{1}}}\ket{\lambda_{N-2}}+\frac{\xi^{N-1}}{\sqrt{\lambda_{N-1}\lambda_{N-2}...\lambda_{1}}}\ket{\lambda_{N-1}})
=λ0|𝐀|λ0+λ1|𝐀|λ1++λN1|𝐀|λN1\displaystyle=\bra{\lambda_{0}}{{\bf A}}\ket{\lambda_{0}}+\bra{\lambda_{1}}{{\bf A}}\ket{\lambda_{1}}+...+\bra{\lambda_{N-1}}{{\bf A}}\ket{\lambda_{N-1}}\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\> (118)

where we have used the regularized integrals from Equation (29).

The action integral is, for a “Hamiltonian” kBTϵkBTαα=𝒦ϵαα\frac{\mathcal{H}}{k_{B}T}\equiv\frac{\epsilon}{k_{B}T}\alpha^{\dagger}\alpha={\mathcal{K}}\epsilon\alpha^{\dagger}\alpha,

𝒵=Tr(e𝒦ϵαα)=𝑑ξ𝑑ξ1ξ|ξ𝒞01ξ|e𝒦ϵαα|ξ\displaystyle\mathcal{Z}=Tr\left(e^{-{\mathcal{K}}\epsilon\alpha^{\dagger}\alpha}\right)=\int d\xi^{\dagger}d\xi\frac{1}{\innerproduct{\xi}{\xi}}\bra{{\cal C}_{0}^{-1}\xi}\>e^{-{\mathcal{K}}\epsilon\alpha^{\dagger}\alpha}\ket{\xi} (119)

The bracketed term can be “split” into sub-integrals using the identity operator from Equation (40). We define ξN=𝒞01ξ0\xi_{N}={\cal C}_{0}^{-1}\xi_{0} and δτ=𝒦N\delta\tau=\frac{{\mathcal{K}}}{N}.

𝒞01ξ|e𝒦ϵαα|ξ=ξN|eδτϵαα(j=1j=N1𝑑ξj𝑑ξj1ξj|ξj|𝒞01ξjξj|eδτϵαα)|ξ0\displaystyle\bra{{\cal C}_{0}^{-1}\>\xi}\>e^{-{\mathcal{K}}\epsilon\alpha^{\dagger}\alpha}\ket{\xi}=\bra{\xi_{N}}e^{-\delta\tau\>\epsilon\>\alpha^{\dagger}\alpha}\bigg{(}\prod_{j=1}^{j=N-1}\int d\xi^{\dagger}_{j}d\xi_{j}\frac{1}{\innerproduct{\xi_{j}}{\xi_{j}}}\ket{{\cal C}_{0}^{-1}\>\xi_{j}^{\dagger}}\bra{\xi_{j}^{\dagger}}e^{-\delta\tau\>\epsilon\>\alpha^{\dagger}\alpha}\bigg{)}\ket{\xi_{0}}
=j=1j=N𝑑ξj𝑑ξjexp(ln(ξj|ξj)+ln(ξj|𝒞01ξj1))eδτϵξj𝒞01ξj1\displaystyle=\prod_{j=1}^{j=N}\int d\xi^{\dagger}_{j}d\xi_{j}\exp{-\ln{\bra{\xi_{j}}\ket{\xi_{j}}}+\ln{\bra{\xi_{j}^{\dagger}}\ket{{\cal C}_{0}^{-1}\>\xi_{j-1}^{\dagger}}}}e^{-\delta\tau\epsilon\>\xi_{j}\>{\cal C}_{0}^{-1}\>\xi^{\dagger}_{j-1}}\>\>\>\> (120)

Using the scalar products as defined in Equation (28) and expanding the logarithm to lowest order,

ξj|ξj1+ξjξj\displaystyle\bra{\xi_{j}}\ket{\xi_{j}}\approx 1+\xi_{j}^{\dagger}\xi_{j}
ln(ξj|ξj)ξjξj\displaystyle\rightarrow\ln{\bra{\xi_{j}}\ket{\xi_{j}}}\approx\xi_{j}^{\dagger}\xi_{j}
ξj|𝒞01ξj11+ξj𝒞01ξj11+ξj1ξj\displaystyle\bra{\xi_{j}^{\dagger}}\ket{{\cal C}_{0}^{-1}\>\xi_{j-1}^{\dagger}}\approx 1+\>\xi_{j}\>{\cal C}_{0}^{-1}\>\xi_{j-1}^{\dagger}\approx 1+\xi_{j-1}^{\dagger}\>\xi_{j}
ln(ξj|𝒞01ξj1)ξj1ξj\displaystyle\rightarrow\ln{\bra{\xi_{j}^{\dagger}}\ket{{\cal C}_{0}^{-1}\>\xi_{j-1}^{\dagger}}}\approx\xi_{j-1}^{\dagger}\xi_{j}
δτϵξj𝒞01ξj1=δτϵξj1ξj\displaystyle\delta\tau\epsilon\>\xi_{j}\>{\cal C}_{0}^{-1}\>\xi^{\dagger}_{j-1}=\delta\tau\epsilon\>\xi^{\dagger}_{j-1}\xi_{j} (121)

the integral reduces to

𝒞01ξ|eδτϵβα|ξ=j=ij=N𝑑ξj𝑑ξjexp((ξjξj1)ξjδτϵξj1ξj)\displaystyle\bra{{\cal C}_{0}^{-1}\>\xi}\>e^{-\delta\tau\epsilon\beta\alpha}\ket{\xi}=\prod_{j=i}^{j=N}\int d\xi^{\dagger}_{j}d\xi_{j}\exp{-(\xi_{j}^{\dagger}-\xi_{j-1}^{\dagger})\xi_{j}-\delta\tau\>\epsilon\>\xi^{\dagger}_{j-1}\xi_{j}}
DξDξexp(𝑑τξ(τ+ϵ)ξ)\displaystyle\Rightarrow\int D\xi^{\dagger}D\xi\exp{-\int d\tau\>\xi^{\dagger}(-\partial_{\tau}+\epsilon)\xi} (122)

with boundary conditions for ξ(τ),τ(0,𝒦)\xi(\tau),\tau\in(0,{\cal K}) appropriate to ξN=𝒞01ξ0\xi_{N}={\cal C}_{0}^{-1}\xi_{0}.

V.1 Periodicity of ξ\xi

To define boundary conditions, we use the Tr(𝒞01)Tr({\cal C}_{0}^{-1}) as the boundary condition. This is consistent with ξ0ξN=ξ0\xi_{0}\rightarrow\xi_{N}=\xi_{0} for bosons and ξ0ξN=ξ0\xi_{0}\rightarrow\xi_{N}=-\xi_{0} for fermions Murayama1 . Hence, defining v=ilogTr(𝒞01)v=i\log Tr({\cal C}_{0}^{-1}),

ξN=ξ0ei(ilogTr(𝒞01))=ξ0eiv\displaystyle\xi_{N}=\xi_{0}\>e^{i(i\log Tr({\cal C}_{0}^{-1}))}=\xi_{0}\>e^{iv} (123)

We generalize

ξ(τ)=q=q=+ζqei(2qπv)𝒦τ\displaystyle\rightarrow\xi(\tau)=\sum_{q=-\infty}^{q=+\infty}\zeta_{q}e^{i\frac{(2q\pi-v)}{\mathcal{K}}\tau} (124)

This leads to the integral for finite-NN, using the rules we postulated before in Equations (29) and (30)

𝒵=q=q=𝑑ζq𝑑ζqexp(ζq(i(2qvπ)π𝒦+ϵ)ζq)\displaystyle\mathcal{Z}=\prod_{q=-\infty}^{q=\infty}\int d\zeta^{\dagger}_{q}d\zeta_{q}\exp{-\zeta_{q}^{\dagger}\>\left(-i\frac{(2q-\frac{v}{\pi})\pi}{\mathcal{K}}+\epsilon\right)\>\zeta_{q}}
q=q=(i(2q1)π(vππ)π𝒦+ϵ)N1=q=0q=[1+(ϵ+ivπ𝒦(2q1)π𝒦)2](N1)\displaystyle\propto\prod_{q=-\infty}^{q=\infty}\left(-i\frac{(2q-1)\pi-(\frac{v-\pi}{\pi})\pi}{\mathcal{K}}+\epsilon\right)^{N-1}=\prod_{q=0}^{q=\infty}\bigg{[}1+\left(\frac{\epsilon+i\frac{v-\pi}{\mathcal{K}}}{\frac{(2q-1)\pi}{\mathcal{K}}}\right)^{2}\bigg{]}^{(N-1)}
=q=0q=[1+(𝒦ϵ+i(vπ)(2q1)π)2](N1)\displaystyle=\prod_{q=0}^{q=\infty}\bigg{[}1+\left(\frac{{\mathcal{K}}\epsilon+i(v-\pi)}{(2q-1)\pi}\right)^{2}\bigg{]}^{(N-1)}
cosh(𝒦ϵ+i(vπ)2)N1\displaystyle\propto\cosh\left(\frac{{\mathcal{K}}\epsilon+i(v-\pi)}{2}\right)^{N-1}

In the above, we use the relations (aa is real in the below and the third equation is a general version of the second)

q=1q=(1+x2q2)=sinhπx2πx2\displaystyle\prod_{q=1}^{q=\infty}\left(1+\frac{x^{2}}{q^{2}}\right)=\frac{\sinh\frac{\pi x}{2}}{\frac{\pi x}{2}}
q=1q=(1+x2(2q1)2)=coshπx2\displaystyle\prod_{q=1}^{q=\infty}\left(1+\frac{x^{2}}{(2q-1)^{2}}\right)=\cosh\frac{\pi x}{2}
q=1q=(1+x2(2qa)2)=Γ(1a2)Γ(aix+22)Γ(a+ix+22)\displaystyle\prod_{q=1}^{q=\infty}\left(1+\frac{x^{2}}{(2q-a)^{2}}\right)=\frac{\Gamma(1-\frac{a}{2})}{\Gamma(\frac{-a-ix+2}{2})\>\Gamma(\frac{-a+ix+2}{2})}
cosh(a+ib)+cosh(aib)=cosh(a)cos(b)\displaystyle\cosh(a+ib)+\cosh(a-ib)=\cosh(a)\cos(b) (126)

Hence, with θ=π,N=2\theta=\pi,N=2, i.e., for fermions, we get 𝒵cosh(𝒦ϵ2)\mathcal{Z}\propto\cosh(\frac{{\mathcal{K}}\epsilon}{2}). Separately, when θ=0,N\theta=0,N\rightarrow\infty, i.e., for bosons, the above formula reduces to the expression for bosons, i.e., 𝒵1sinh(𝒦ϵ2)\mathcal{Z}\propto\frac{1}{\sinh(\frac{{\mathcal{K}}\epsilon}{2})}

VI Differentiation of generalized Grassmann variables

We wish to replicate the operator θ\theta-commutator (α,αT)θ=1\bigg{(}\alpha,\alpha^{T}\bigg{)}_{\theta}=1 with coherent state variables. Realizing that ξ\xi’s have to be treated as matrices and comparing the situation with Equation (1) and (10), we impose,

ξ(ξf)=f+z1ξξf\displaystyle\frac{\partial}{\partial\xi}\bigg{(}\xi f\bigg{)}=f+z^{-1}\>\xi\frac{\partial}{\partial\xi}f
ξT(ξTf)=f+zξTξTf\displaystyle\frac{\partial}{\partial\xi^{T}}\bigg{(}\xi^{T}f\bigg{)}=f+z\>\xi^{T}\frac{\partial}{\partial\xi^{T}}f (127)

so that there is a zz or z1z^{-1} that accompanies switching the derivative and the variable. This is consistent with the bosonic and fermionic case and permits us to deduce the uncertainty relation for the coherent variable and its conjugate momentum, i.e.,

(ξ,ξ)θf(θ)=ξ(ξf)z1ξξf=f\displaystyle\bigg{(}\frac{\partial}{\partial\xi},\xi\bigg{)}_{-\theta}f(\theta)=\frac{\partial}{\partial\xi}(\xi f)-z^{-1}\>\xi\frac{\partial}{\partial\xi}f=f
(ξT,ξT)θf(θ)=ξT(ξTf)zξTξTf=f\displaystyle\bigg{(}\frac{\partial}{\partial\xi^{T}},\xi^{T}\bigg{)}_{\theta}f(\theta)=\frac{\partial}{\partial\xi^{T}}(\xi^{T}f)-z\>\xi^{T}\frac{\partial}{\partial\xi^{T}}f=f (128)

An immediate consequence is

ξξn=(1+z1+z2++z(n1))ξn1\displaystyle\frac{\partial}{\partial\xi}\xi^{n}=\left(1+z^{-1}+z^{-2}+...+z^{-(n-1)}\right)\xi^{n-1}
ξT(ξT)n=(1+z1+z2++zn1)(ξT)n1\displaystyle\frac{\partial}{\partial\xi^{T}}(\xi^{T})^{n}=\left(1+z^{1}+z^{2}+...+z^{n-1}\right)(\xi^{T})^{n-1} (129)

The above equation is consistent on both sides if we were to set n=Nn=N, for ξN=(ξT)N=0\xi^{N}=(\xi^{T})^{N}=0, as well as 1+z1+z2++zN1=01+z^{1}+z^{2}+...+z^{N-1}=0 and 1+z1+z2++z(N1)=01+z^{-1}+z^{-2}+...+z^{-(N-1)}=0.

We are going to assume that ξ\xi and ξT,ξ\xi^{T},\xi^{\dagger} are all independent of each other. By demanding consistency as in

ξξξT=ξT=ξzξTξ\displaystyle\frac{\partial}{\partial\xi}\xi\xi^{T}=\xi^{T}=\frac{\partial}{\partial\xi}z\xi^{T}\xi
ξTξTξ=ξ=ξTz1ξξT\displaystyle\frac{\partial}{\partial\xi^{T}}\xi^{T}\xi=\xi=\frac{\partial}{\partial\xi^{T}}z^{-1}\xi\xi^{T} (130)

we deduce the rules for switching the partial derivative and the transposed variable,

ξ(ξTf)=z1ξTξf\displaystyle\frac{\partial}{\partial\xi}(\xi^{T}f)=z^{-1}\xi^{T}\frac{\partial}{\partial\xi}f
ξT(ξf)=zξξTf\displaystyle\frac{\partial}{\partial\xi^{T}}(\xi f)=z\xi\frac{\partial}{\partial\xi^{T}}f (131)

Also,

ξeξξ=ξeξξ\displaystyle\frac{\partial}{\partial\xi}e^{-\xi\xi^{\dagger}}=-\xi^{\dagger}e^{-\xi\xi^{\dagger}}
ξeξξ=𝒞0ξeξξ\displaystyle\frac{\partial}{\partial\xi^{\dagger}}e^{-\xi\xi^{\dagger}}=-{\cal C}_{0}\>\xi e^{-\xi\xi^{\dagger}} (132)

In addition, by considering switching the order of partial derivatives, we get

ξTξ=z1ξξT\displaystyle\frac{\partial}{\xi^{T}}\frac{\partial}{\partial\xi}=z^{-1}\frac{\partial}{\partial\xi}\frac{\partial}{\xi^{T}} (133)

which is derived from

ξTξξξT=1=z1ξξTzξTξ\displaystyle\frac{\partial}{\xi^{T}}\frac{\partial}{\partial\xi}\>\xi\xi^{T}=1=z^{-1}\frac{\partial}{\partial\xi}\frac{\partial}{\xi^{T}}\>z\>\xi^{T}\xi (134)

VII 2-d and 1-d harmonic oscillator: Generalized Hermite polynomials

Let’s solve for the two-dimensional oscillator first. The operators α,αT\alpha,\alpha^{T} are the usual annihilation and creation operators. Let’s assume 𝒜,Γ,Υ{\cal A},\Gamma,\Upsilon are all complex numbers that commute with the ξ,ξT\xi,\xi^{T}.

α=𝒜2(ξ+ΓξT)\displaystyle\alpha=\frac{{\cal A}}{\sqrt{2}}\left(\xi+\Gamma\frac{\partial}{\partial\xi^{T}}\right)
αT=𝒜2(ξT+Υξ)\displaystyle\alpha^{T}=\frac{{\cal A}}{\sqrt{2}}\left(\xi^{T}+\Upsilon\frac{\partial}{\partial\xi}\right) (135)

If we want this to be consistent with

ααTzαTα=1\displaystyle\alpha\alpha^{T}-z\alpha^{T}\alpha=1 (136)

we derive the simplest solution, that matches the conditions for the case of fermions as well as bosons, i.e.,

Γ=1\displaystyle\Gamma=1
Υ=1z\displaystyle\Upsilon=-\frac{1}{z}
𝒜=1\displaystyle{\cal A}=1 (137)

i.e.,

α=12(ξ+ξT)\displaystyle\alpha=\frac{1}{\sqrt{2}}\left(\xi+\frac{\partial}{\partial\xi^{T}}\right)
αT=12(ξTz1ξ)\displaystyle\alpha^{T}=\frac{1}{\sqrt{2}}\left(\xi^{T}-z^{-1}\frac{\partial}{\partial\xi}\right) (138)

The ground state wave-function f0f_{0} is found from αf0=0\alpha f_{0}=0, which leads to

12(ξ+ξT)f0=0\displaystyle\frac{1}{\sqrt{2}}\left(\xi+\frac{\partial}{\partial\xi^{T}}\right)f_{0}=0
f0(ξT)meξTξ=(ξT)mez1ξξT\displaystyle\rightarrow f_{0}\sim(\xi_{T})^{m}e^{-\xi^{T}\xi}=(\xi_{T})^{m}e^{-z^{-1}\xi\xi^{T}} (139)

Note that upon expanding the exponential multiplying by the polynomial in ξ)T\xi)T, we’d keep terms up to ξN1,(ξT)N1\xi^{N-1},(\xi^{T})^{N-1} as higher powers are 0. We can construct higher wavefunctions using the creation operator, f1=αTf0f_{1}=\alpha^{T}f_{0} etc.

To obtain the wave-function for the 1-d harmonic oscillator, it is reasonable to assume (in line with the method one uses with the bosonic caseWess ) that α,αT\alpha,\alpha^{T} are real. In addition, to not allow powers of ξT\xi^{T} in the ground-state function, we set ξ=ξT\xi=\xi^{T} . We then deduce

α=12(ξ+ξ)\displaystyle\alpha=\frac{1}{\sqrt{2}}\left(\xi+\frac{\partial}{\partial\xi}\right)
αT=12(ξcosθξ)\displaystyle\alpha^{T}=\frac{1}{\sqrt{2}}\left(\xi-\cos\theta\>\frac{\partial}{\partial\xi}\right) (140)

which yields, when one carries out the above construction, terms that reduce to Hermite polynomials (albeit with series and exponentials terminated at xN1,yN1x^{N-1},y^{N-1}), i.e.,

f0Ceξ22\displaystyle f_{0}\sim Ce^{-\frac{\xi^{2}}{2}}
f112C(1+cosθ)ξeξ22\displaystyle f_{1}\sim\frac{1}{\sqrt{2}}C(1+\cos\theta)\xi e^{-\frac{\xi^{2}}{2}}
f212C(1+cosθ)(ξ2(1+cosθ)cosθ)eξ22\displaystyle f_{2}\sim\frac{1}{\sqrt{2}}C(1+\cos\theta)(\xi^{2}(1+\cos\theta)-\cos\theta)e^{-\frac{\xi^{2}}{2}} (141)

We note that this construction does not work for θ=π\theta=\pi, the α\alpha’s cannot be real.

VIII Wave-function for two anyons

The excitations described in this paper represent one of the two ways two anyons can be braided amongst each other. While we have yet to construct a description for anyons here, we could consider start by studying “overons” (as opposed to “underons” which have the complex conjugate eigenvalues).

For two dimensions, we had, for the ground state for overons (as well as for underons)

f0=C(ξT)meξTξ\displaystyle f_{0}=C\>(\xi^{T})^{m}\>e^{-\xi^{T}\xi} (142)

which are the usual holomorphic functions.

For two excitations, the wave-function needs to possess the proper symmetry upon exchange, hence would be

f(ξ1,ξ1T,ξ2,ξ2T)=C(ξ1T+zξ2T)Neξ1Tξ1ξ2Tξ2\displaystyle f(\xi_{1},\xi_{1}^{T},\xi_{2},\xi_{2}^{T})=C(\xi_{1}^{T}+z\xi_{2}^{T})^{N}e^{-\xi_{1}^{T}\xi_{1}-\xi_{2}^{T}\xi_{2}} (143)

This works as the exchange 121\leftrightarrow 2 causes (ξ1T+zξ2T)N(\xi_{1}^{T}+z\xi_{2}^{T})^{N} to go to

(ξ1T+zξ2T)(ξ2T+zξ1T)N=(ξ1T+z1ξ2T)\displaystyle(\xi_{1}^{T}+z\xi_{2}^{T})\rightarrow(\xi_{2}^{T}+z\xi_{1}^{T})^{N}=(\xi_{1}^{T}+z^{-1}\xi_{2}^{T}) (144)

which produces a wave-function for the excitation with the opposite exchange characteristic, i.e., for “underons”. While this would not be a possible symmetry for “overons”, it could represent an appropriate anyon wave-function.

However, as can be quickly checked, the overlap of this function with the Laughlin-like Laughlin alternative

f(ξ1,ξ1T,ξ2,ξ2T)=C(ξ1Tξ2T)Neξ1Tξ1ξ2Tξ2\displaystyle f_{\cal L}(\xi_{1},\xi_{1}^{T},\xi_{2},\xi_{2}^{T})=C(\xi_{1}^{T}-\xi_{2}^{T})^{N}e^{-\xi_{1}^{T}\xi_{1}-\xi_{2}^{T}\xi_{2}} (145)

is non-zero only for odd NN, as only terms with both ξTξ\xi^{T}\xi raised to powers gives non-zero results.

This is clear, as looking at individual non-zero terms in the overlap integral, i.e.,

𝑑ξ1T𝑑ξ1𝑑ξ2T𝑑ξ2(ξ1T+zξ2T)N(ξ1Tξ2T)Ne2ξ1Tξ12ξ2Tξ2\displaystyle\int d\xi_{1}^{T}d\xi_{1}d\xi_{2}^{T}d\xi_{2}\>(\xi_{1}^{T}+z\xi_{2}^{T})^{N}\>(\xi_{1}^{T}-\xi_{2}^{T})^{N}\>e^{-2\xi_{1}^{T}\xi_{1}-2\xi_{2}^{T}\xi_{2}}
m=0N1(ξ1T)m(zξ2T)Nm(ξ1)m(ξ2)Nm(ξ1Tξ1)Nm(ξ2Tξ2)m\displaystyle\propto\sum_{m=0}^{N-1}(\xi_{1}^{T})^{m}(z\xi_{2}^{T})^{N-m}(\xi_{1})^{m}(-\xi_{2})^{N-m}\>(-\xi^{T}_{1}\xi_{1})^{N-m}(-\xi_{2}^{T}\xi_{2})^{m}
=m=0N1(z)m=1(1)N11z\displaystyle=\sum_{m=0}^{N-1}(-z)^{-m}=\frac{1-(-1)^{N}}{1-\frac{1}{z}} (146)

which shows that the overlap integral is 0 for even-NN and proportional to (=211z=\frac{2}{1-\frac{1}{z}}) for odd-NN. The overlap is of order unity for small odd-NN, which explains why the function works so well, even for particles with fractional statistics.

IX Remarks on the propagation of Anyons

In the usual field theory of fermions Banks ; QFT , the evolution of a two-particle state is expressed as a perturbative series, starting with the two “bare” propagators followed by successively more complex interactions, involving vertices of the interactions and loops between such vertices.

Let’s say we start with two identical fermions (11 and 22), treated as a product of fairly well-separated wave-functions, at points AA or BB. Starting at these separated spots, they can be end up at two widely-separated spots with in all possible ways - either through direct propagation, or with crossed propagation, i.e.,

ψTotal(x1,x2)=ψA(x1)ψB(x2)ψA(x2)ψB(x1)\displaystyle\psi_{Total}(x_{1},x_{2})=\psi_{A}(x_{1})\psi_{B}(x_{2})-\psi_{A}(x_{2})\psi_{B}(x_{1}) (147)

Usually, the wave-function with exchanged coordinates (the second term) is extremely small, so we can approximate the propagation neglecting the crossed term.

Figure 4: Two anyon propagation
Refer to caption
Refer to caption

The propagator for two anyons, however, has an interesting twist. The propagation of “bare” anyons is connected to the value of θ\theta. Two anyons can propagate, as in Fig. 2 while winding around each other 0,2,4,… times (since the final result would then be indistinguishable from no winding) and we need to sum over all these possible alternatives. Suppose θ=2Mπ2Q+1\theta=\frac{2M\pi}{2Q+1} where QQ is an integer and the integers M, 2Q+1M,\>2Q+1 are co-prime. Each “winding” produces a multiplicative factor of σ^=eiθ{\hat{\sigma}}=e^{i\theta} into the amplitude for the two-anyon propagator. Since we need to sum over all the separate ways this can happen, we have to restrict the number of windings to be less than 2Q+12Q+1, when we revert to zero windings. The following factors are useful to define.

a1=1+σ^2+σ^4++σ^2Qeiθ22cosθ2\displaystyle a_{1}=1+{\hat{\sigma}}^{2}+{\hat{\sigma}}^{4}+...+{\hat{\sigma}}^{2Q}\equiv\frac{e^{-i\frac{\theta}{2}}}{2\cos\frac{\theta}{2}}
b1=1+σ^2+σ^4++σ^2Qeiθ22cosθ2\displaystyle b_{1}=1+{\hat{\sigma}}^{-2}+{\hat{\sigma}}^{-4}+...+{\hat{\sigma}}^{-2Q}\equiv\frac{e^{i\frac{\theta}{2}}}{2\cos\frac{\theta}{2}}
a2=σ^+σ^3++σ^2Q+1eiθ22cosθ2\displaystyle a_{2}={\hat{\sigma}}+{\hat{\sigma}}^{3}+...+{\hat{\sigma}}^{2Q+1}\equiv\frac{e^{i\frac{\theta}{2}}}{2\cos\frac{\theta}{2}}
b2=σ^1+σ^3++σ^(2Q+1)eiθ22cosθ2\displaystyle b_{2}={\hat{\sigma}}^{-1}+{\hat{\sigma}}^{-3}+...+{\hat{\sigma}}^{-(2Q+1)}\equiv\frac{e^{-i\frac{\theta}{2}}}{2\cos\frac{\theta}{2}}

so that the amplitude for the propagation of two anyons is proportional to S1+S2=1S_{1}+S_{2}=1 since

S1=a1+b11\displaystyle S_{1}=a_{1}+b_{1}\equiv 1
S2=a2+b21\displaystyle S_{2}=a_{2}+b_{2}\equiv 1 (148)
Figure 5: Odd and Even Roots of Unity
Refer to caption
Refer to caption

On the other hand, if θ=2Mπ2Q\theta=\frac{2M\pi}{2Q}, with the caveat |Q|>1\absolutevalue{Q}>1 and M, 2QM,\>2Q co-prime, then the corresponding sums are

p1=1+σ^2+σ^4++σ^2(Q1)0\displaystyle p_{1}=1+{\hat{\sigma}}^{2}+{\hat{\sigma}}^{4}+...+{\hat{\sigma}}^{2(Q-1)}\equiv 0
q1=1+σ^2+σ^4++σ^2(Q1)0\displaystyle q_{1}=1+{\hat{\sigma}}^{-2}+{\hat{\sigma}}^{-4}+...+{\hat{\sigma}}^{-2(Q-1)}\equiv 0
p2=σ^+σ^3++σ^2Q10\displaystyle p_{2}={\hat{\sigma}}+{\hat{\sigma}}^{3}+...+{\hat{\sigma}}^{2Q-1}\equiv 0
q2=σ^1+σ^3++σ^(2Q1)0\displaystyle q_{2}={\hat{\sigma}}^{-1}+{\hat{\sigma}}^{-3}+...+{\hat{\sigma}}^{-(2Q-1)}\equiv 0

so that the amplitude for the propagation is 12(R1+R2)=0\frac{1}{2}(R_{1}+R_{2})=0 below since

R1=p1+q10\displaystyle R_{1}=p_{1}+q_{1}\equiv 0
R2=p2+q20\displaystyle R_{2}=p_{2}+q_{2}\equiv 0 (149)

The exclusion of Q=1Q=1 (in the ‘even’ case above) is also easy to see (this explicitly means two fermions can indeed propagate). Note that if Q=1Q=1, the individual terms p1,q1,p2,q2p_{1},q_{1},p_{2},q_{2} are all 11 or 1-1. There are no cancellations.

Anyons with even-denominator θ\theta cannot propagate freely, purely from geometrical considerations, while those with odd-denominator θ\theta can.

This result might seem surprising, but it is clear from studying the roots of unity on the complex plane. Consider Fig. 3, where the odd and even roots of unity are displayed for the case of 3 and 4, respectively. It is clear, for instance, that the sum 1+e2iθ1+e^{2i\theta} is non-zero in the odd-case, while the sum 1+e2iθ1+e^{2i\theta} is zero in the even case.

The factors S1,S2,R1,R2S_{1},S_{2},R_{1},R_{2} multiply the total amplitudes, assuming there are no energetic consequences within the action to meandering paths that wind around each other multiple number of times. It would, therefore, not be surprising that multiple anyon propagation is suppressed with even-denominator theta. It is, however, possible that higher order interaction terms will allow propagation for even-denominator-theta anyons. This will have physical consequences for transport and localization of states with odd and even denominator θ\theta.

This behavior (for odd and even-denominator θ\theta) is quite robust to the addition of a small cost for extra windings. Suppose the anyons encountered a cost, a factor of h=1δh=1-\delta, small δ\delta cost, for each extra winding around the other anyon. Then the corresponding sums would become, for even-denominator θ=2Mπ2Q\theta=\frac{2M\pi}{2Q},

p^1=1+h2σ^2+h4σ^4++(h2)Q1σ^2(Q1)δ2Q1e2iθ\displaystyle\hat{p}_{1}=1+h^{2}{\hat{\sigma}}^{2}+h^{4}{\hat{\sigma}}^{4}+...+(h^{2})^{Q-1}{\hat{\sigma}}^{2(Q-1)}\approx\delta\frac{2Q}{1-e^{2i\theta}}
q^1=1+h2σ^2+h4σ^4++(h2)Q1σ^2(Q1)δ2Q1e2iθ\displaystyle\hat{q}_{1}=1+h^{2}{\hat{\sigma}}^{-2}+h^{4}{\hat{\sigma}}^{-4}+...+(h^{2})^{Q-1}{\hat{\sigma}}^{-2(Q-1)}\approx\delta\frac{2Q}{1-e^{-2i\theta}}
p^2=hσ^+h3σ^3++h2Q1σ^2Q1heiθδ2Q1e2iθ\displaystyle\hat{p}_{2}=h{\hat{\sigma}}+h^{3}{\hat{\sigma}}^{3}+...+h^{2Q-1}{\hat{\sigma}}^{2Q-1}\approx he^{i\theta}\delta\frac{2Q}{1-e^{2i\theta}}
q^2=hσ^1+h3σ^3++h2Q1σ^(2Q1)heiθδ2Q1e2iθ\displaystyle\hat{q}_{2}=h{\hat{\sigma}}^{-1}+h^{3}{\hat{\sigma}}^{-3}+...+h^{2Q-1}{\hat{\sigma}}^{-(2Q-1)}\approx he^{-i\theta}\delta\frac{2Q}{1-e^{-2i\theta}}
R^1=p^1+q^1=2Qδ\displaystyle\hat{R}_{1}=\hat{p}_{1}+\hat{q}_{1}=2Q\delta
R^2=p^2+q^2=0(toorderδ)\displaystyle\hat{R}_{2}=\hat{p}_{2}+\hat{q}_{2}=0\>(\>to\>order\>\delta) (150)

while the odd-denominator θ=2π2Q+1\theta=\frac{2\pi}{2Q+1} results are of 1\sim 1 to the same order. The suppression is, therefore, robust for small additional cost for anyons “winding” around each other.

X Conclusions

We have studied the arithmetic and calculus of overons, excitations under generalized statistics interpolating between fermions and bosons. We have described the fuzzy “pancake” surface that best describes the eigenvalue surface for the algebra. Further, the calculus of coherent state variables is studied, as is the partition function for these states. We then proceed to study generalizations of the Hermite polynomials.

Using the results, we have studied some consequences for the field theory of anyons that immediately result from the calculus of coherent state variables as well as from the geometrical interpretation. These demonstrate the appropriateness of the Laughlin wave-function to describe 2-anyon states. In addition, there are rather simple geometrical reasons why even-denominator θ\theta anyons cannot propagate freely and can only do so in the presence of anyon-anyon interactions.

XI Acknowledgments

SR acknowledges the hospitality and intellectual stimulation of the Rutgers Department of Physics & Astronomy and the NHETC at Rutgers. Much of this work benefited from very useful advice and suggestions from Professor Scott Thomas. He also acknowledges the collaborative atmosphere provided at the ITP, Santa Barbara.

XII APPENDIX 1: A dynamical system analog for the eigenvalue spectrum

XII.1 Boson starting point

Consider the problem of a boson, represented as a scalar field, defined on a circle (S1)(S^{1}) around another boson. The circle is discretized into NN points, labelled 0,1,,N10,1,...,N-1. The lattice constant (between the discrete points) is a=2πN=θa=\frac{2\pi}{N}=\theta. The potential energy part of the Hamiltonian, after partial integration, leads to

Pboson=j=0N1𝒦ϕ(j)2(ϕ(j+1)ϕ(j)aϕ(j)ϕ(j1)a)\displaystyle{\cal H}_{P}^{boson}=-\sum_{j=0}^{N-1}{\cal K}\frac{\phi(j)}{2}\left(\frac{\phi(j+1)-\phi(j)}{a}-\frac{\phi(j)-\phi(j-1)}{a}\right) (151)

Transform to Fourier coordinates

ϕj=1Nk=0N1ϕ~(k)ei2πNakja\displaystyle\phi_{j}=\frac{1}{\sqrt{N}}\sum_{k=0}^{N-1}{\tilde{\phi}}(k)\>e^{i\frac{2\pi}{Na}kja}
Pboson=𝒦ak=0N1ϕ~(k)ϕ~(k)(1cos(2πNaka))\displaystyle\rightarrow{\cal H}_{P}^{boson}=\frac{{\cal K}}{a}\sum_{k=0}^{N-1}{\tilde{\phi}}(k){\tilde{\phi}}(-k)\>(1-\cos(\frac{2\pi}{Na}ka))
=2𝒦ak=0N1ϕ~(k)ϕ~(k)sin2(kθ2)\displaystyle=2\frac{{\cal K}}{a}\sum_{k=0}^{N-1}{\tilde{\phi}}^{*}(k){\tilde{\phi}}(k)\>\sin^{2}(\frac{k\theta}{2}) (152)

XII.2 Fermion starting point

Start with a fermion, represented by a complex field defined on the same circle. The Hamiltonian would have a first-order derivative, as below

Pfermion=ij=0N1𝒦2ϕ(j)(ϕ(j)ϕ(j1)a)+c.c.\displaystyle{\cal H}_{P}^{fermion}=-i\sum_{j=0}^{N-1}\frac{{\cal K}}{2}\phi^{*}(j)\left(\frac{\phi(j)-\phi(j-1)}{a}\right)+c.c. (153)

Again, writing this in Fourier space, using a=2πNa=\frac{2\pi}{N},

Pfermion=𝒦ak=0N1ϕ~(k)ϕ~(k)eika2sin(ka2)+c.c.=2𝒦ak=0N1ϕ~(k)ϕ~(k)cos(k2πN2)sin(k2πN2)\displaystyle{\cal H}_{P}^{fermion}=\frac{{\cal K}}{a}\sum_{k=0}^{N-1}{\tilde{\phi}}^{*}(k){\tilde{\phi}}(k)e^{\frac{-ika}{2}}\sin(\frac{ka}{2})+c.c.=\frac{2{\cal K}}{a}\sum_{k=0}^{N-1}{\tilde{\phi}}^{*}(k){\tilde{\phi}}(k)\cos(\frac{k\frac{2\pi}{N}}{2})\sin(\frac{k\frac{2\pi}{N}}{2})
=𝒦ak=0N1ϕ~(k)ϕ~(k)sin(kθ)\displaystyle=\frac{{\cal K}}{a}\sum_{k=0}^{N-1}{\tilde{\phi}}^{*}(k){\tilde{\phi}}(k)\sin(k\theta)\>\>\>\>\>\>\>\>\>\>\>\> (154)

Fermion doubling - this would go away if we set a=πNa=\frac{\pi}{N}.

XII.3 A composite particle starting point

Composing the above Hamiltonians, choosing to appropriately interpolate between the boson (θ=0)(\theta=0) and fermion (θ=π)(\theta=\pi) cases, we write the discretized version of the fractional (ν\nu) derivative as follows (assume wrap-around coordinatization for a circle)

Pcomposite=(i)ν𝒦aνj=0N1ϕ(j)(ϕ(j)+(1)Γ(ν+1)1!Γ(ν)ϕ(j1)+(1)2Γ(ν+1)2!Γ(ν1)ϕ(j2)\displaystyle{\cal H}_{P}^{composite}=(-i)^{\nu}\frac{{\cal K}}{a^{\nu}}\sum_{j=0}^{N-1}\phi^{*}(j)\bigg{(}\phi(j)+(-1)\frac{\Gamma(\nu+1)}{1!\Gamma(\nu)}\phi(j-1)+(-1)^{2}\frac{\Gamma(\nu+1)}{2!\Gamma(\nu-1)}\phi(j-2)
++(1)N1Γ(ν+1)(N1)!Γ(νN+1)ϕ(j(N1)))+c.c.\displaystyle+...+(-1)^{N-1}\frac{\Gamma(\nu+1)}{(N-1)!\Gamma(\nu-N+1)}\phi(j-(N-1))\bigg{)}+c.c.\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\> (155)

which, in Fourier space becomes

Pcomposite=(i)ν𝒦aνk=0N1ϕ~(k)ϕ~(k)(1+(1)Γ(ν+1)1!Γ(ν)eika+(1)2Γ(ν+1)2!Γ(ν1)e2ika\displaystyle{\cal H}_{P}^{composite}=(-i)^{\nu}\frac{{\cal K}}{a^{\nu}}\sum_{k=0}^{N-1}{\tilde{\phi}^{*}(k)}{\tilde{\phi}(k)}\bigg{(}1+(-1)\frac{\Gamma(\nu+1)}{1!\Gamma(\nu)}e^{-ika}+(-1)^{2}\frac{\Gamma(\nu+1)}{2!\Gamma(\nu-1)}e^{-2ika}
++(1)N1Γ(ν+1)(N1)!Γ(νN+1)e(N1)ika)+c.c.\displaystyle+...+(-1)^{N-1}\frac{\Gamma(\nu+1)}{(N-1)!\Gamma(\nu-N+1)}e^{-(N-1)ika}\bigg{)}+c.c.\>\>\>\>\>\>
=(i)ν𝒦aνk=0N1ϕ~(k)ϕ~(k)(1eika)ν\displaystyle=(-i)^{\nu}\frac{{\cal K}}{a^{\nu}}\sum_{k=0}^{N-1}{\tilde{\phi}^{*}(k)}{\tilde{\phi}(k)}\bigg{(}1-e^{-ika}\bigg{)}^{\nu}\>\>\>\>\>\>
=4ν𝒦aνcoskaν2(sinka2)ν\displaystyle=4^{\nu}\frac{{\cal K}}{a^{\nu}}\cos\frac{ka\nu}{2}\bigg{(}\sin\frac{ka}{2}\bigg{)}^{\nu}\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\>\> (156)

This is a version of fermion doubling for the composite particles.

References

  • (1) Ramakrishna, Satish Algebra for Fractional Statistics - interpolating from fermions to bosons arxiv:2005.02172, submitted to Nucl. Phys. B
  • (2) Ramakrishna, Satish A generalized q-commutator and other diversions Paper in preparation
  • (3) Chaichian, M., Demichev, A. P. Path Integrals with Generalized Grassmann Variables arxiv:9504016
  • (4) MacMahon, P.A. Combinatory Analysis, Vol. 1 and 2, Cambridge Univ.Press, Cambridge, 1915, reprinted by Chelsea, New York (1955).
  • (5) Fivel, Daniel I. Interpolation between Fermi and Bose Statistics Using Generalized Commutators Phy. Rev. Lett. 65, 3361 (1990)
  • (6) Iida, Shinji, Kuratsuji, Hiroshi Quantum Algebra near q=1 and a Deformed Symplectic Structure Phy. Rev. Lett. 69, 1833 (1992)
  • (7) Greenberg, Oscar W. Interactions of particles having small violations of statistics Physica A 180, 419-427 (1992)
  • (8) Leinaas, J. M., Myrheim, J, Nuovo Cimento B 37 (1977) 1.
  • (9) Rao, Sumathi Anyons: a primer, arxiv.org/abs/hep-th/9209066
  • (10) Khare, Avinash Fractional Statistics And Quantum Theory, World Scientific, 2nded.2^{nd}\>ed. 2005.
  • (11) Boschi-Filho, H. , Farina, C. , de Souza Dutra, A. , The Partition Function for an Anyon-Like Oscillator, arxiv.org/abs/hep-th/9410098
  • (12) Lorek A., Ruffing A., Wess J., A q-deformation of the Harmonic Oscillator Zeitschrift fur Physik C Particles and Fields, June 1997, Volume 74, Issue 2, pp. 369-377 also arxiv:hep-th/960516v1 22 May 1996
  • (13) Sang, Chung Won New q-Deformed Fermionic Oscillator Algebra and Thermodynamics, J. Adv. Physics Vol. 4, pp. 1-4, 2015
  • (14) MacFarlane A. J., On q-analogues of the quantum harmonic oscillator and the quantum group SU(2)qSU(2)_{q} J. Phys. A.: Math. Gen 22 4581
  • (15) Biedenharn L. C., The quantum group SUq(2) and a q-analogue of the boson operators J. Phys. A.: Math. Gen 22 L873
  • (16) Murayama, H. Unpublished Lecture Notes “Statistical Mechanics of Bosons” http://hitoshi.berkeley.edu/221B/bosons.pdf
  • (17) Murayama, H. Unpublished Lecture Notes “Statistical Mechanics of Fermions” http://hitoshi.berkeley.edu/221B/fermions.pdf
  • (18) Lavagno A., Narayana Swamy P. Generalized thermodynamics of q-deformed bosons and fermions arxiv:cond-mat/0111112v1 7 Nov 2000
  • (19) Banks, T. Quantum Mechanics: An Introduction Taylor & Francis (2018)
  • (20) Griffiths, David J. Introduction to Quantum Mechanics Prentics Hall (1994)
  • (21) Laughlin R. B., Anomalous Quantum Hall Effect: An incompressible Quantum Fluid with Fractionally Charged Excitations Phys. Rev. Lett. 50, 18 (1983)
  • (22) Hallnas, Martin The fuzzy sphere, a noncommutative geometry http://courses.theophys.kth.se/SI2350/fuzzy.pdf
  • (23) Oliveira, E. C de, Machado, J. A. T. A Review of Definitions for Fractional Derivatives and Integral Vol. 2014 doi.org/10.1155/2014/238459