1 General Setup and SDE Representation

We consider the d-dimensional Bose-Hubbard model with Hamiltonian

$$\begin{aligned} H= -J{\sum }_{\langle i,j \rangle } {a_{i}^{+}}{a_{j}} + \frac{U}{2} {\sum }_{j} {a_{j}^{+}a_{j}^{+}}a_{j}a_{j} +{\sum }_{j} \epsilon _{j}{a_{j}^{+}}a_{j} \end{aligned}$$
(1.1)

with bosonic annihilation and creation operators \(a_j,a_j^+\) satisfying the commutation relations

$$\begin{aligned}{}[\,a_i\,,a_j^+]= & {} \delta _{i,j} \end{aligned}$$
(1.2)

As usual, \(\langle i,j\rangle \) denotes the sum over nearest neighbors and to be specific, we choose a cubic lattice \(\Gamma \) given by

$$\begin{aligned} j\,=\,(j_1,\cdots ,j_d)\;\in \;\{1,2,\cdots ,L\}^d\;=:\;\Gamma \end{aligned}$$
(1.3)

We find it convenient to work in the Bargmann-Segal representation [1] where the \(a_j,a_j^+\) are realized through the operators

$$\begin{aligned} a_j\;=\; {\partial \over \partial z_j} \;,\;\;\;\; a_j^+\;=\; z_j \end{aligned}$$
(1.4)

which act on the Hilbert space of analytic functions of \(|\Gamma |=L^d\) complex variables

$$\begin{aligned} \mathcal{F}:= & {} \Bigl \{\; f=f(\,\{z_j\}\,)\,:\,\mathbb C^{|\Gamma |}\rightarrow \mathbb C\;\;\textrm{analytic}\;\;\Bigr |\;\;\;\Vert f\Vert ^2_\mathcal{F}=(f,f)_{\mathcal{F}}\,<\,\infty \; \Bigr \} \end{aligned}$$
(1.5)

with scalar product

$$\begin{aligned} (f,g)_\mathcal{F}:= & {} \int _{\mathbb C^{|\Gamma |}=\mathbb R^{2{|\Gamma |}}}\, f(z)\,\overline{g(z)}\; d\mu (z) \end{aligned}$$
(1.6)
$$\begin{aligned} d\mu (z):= & {} \prod _{j} \, e^{-|z_j|^2}\; {d\textrm{Re}z_j \,d\textrm{Im}z_j \over \pi } \end{aligned}$$
(1.7)

In the following, sums \(\,\Sigma _j\,\) or \(\,\Sigma _{i,j}\,\) or products \(\,\mathop \Pi _j\,\) are always meant to be sums and products over all lattice sites if not specified otherwise. That is, we use the notation

$$\begin{aligned} {\sum }_j\cdots:= & {} {\sum }_{j\in \Gamma } \cdots \end{aligned}$$
(1.8)
$$\begin{aligned} \prod _j\cdots:= & {} \prod _{j\in \Gamma } \cdots \end{aligned}$$
(1.9)

and \(\,\Sigma _{i,j}\,:=\,\Sigma _{i,j\in \Gamma }\,\). Actually we can allow for a general hopping matrix which should be real and symmetric,

$$\begin{aligned} \varepsilon:= & {} (\,\varepsilon _{ij}\,)\;\;\in \;\;{\mathbb R}^{{|\Gamma |}\times {|\Gamma |}} \end{aligned}$$
(1.10)

with \(\,\varepsilon _{ij}=\varepsilon _{ji}\,\). With that, the final Hamiltonian, we use a small h instead of a capital H, reads

$$\begin{aligned} h= & {} h_0\,+\,h_\textrm{int} \end{aligned}$$
(1.11)

with a quadratic part

$$\begin{aligned} h_0= & {} {\sum }_{i,j}{\,\varepsilon _{ij}\, z_i\,{\partial \over \partial z_j} } \end{aligned}$$
(1.12)

and a quartic part

$$\begin{aligned} h_\textrm{int}= & {} u\,{\sum }_j \, z_j^2\,{\partial ^2\over \partial z_j^2} \end{aligned}$$
(1.13)

where we substituted the capital U by a small u according to (the capital U’s we use later for a unitary evolution matrix)

$$\begin{aligned} u:= & {} {U\over 2} \end{aligned}$$
(1.14)

For a nearest neighbor hopping J and trapping potentials \(\epsilon _j\) as in (1.1) we have

$$\begin{aligned} \varepsilon _{ij}= & {} {\left\{ \begin{array}{ll} -J &{}\;\textrm{if}\;\;|i-j|=1 \\ +\epsilon _j &{}\;\textrm{if}\;\;i=j \\ 0 &{}\;\mathrm{otherwise\;.} \end{array}\right. } \end{aligned}$$
(1.15)

1.1 Time Evolution of States as Fresnel Expectation Values

The time evolution \(\;e^{-ith}=e^{-it(h_0+h_\textrm{int})}\;\) we calculate through the Trotter formula. That is, we discretize time

$$\begin{aligned} t\;\;=\;\;t_k\;\;=\;\;k\, dt \end{aligned}$$
(1.16)

and write

$$\begin{aligned} e^{-ith}\;\;=\;\;e^{-i\,kdt\,(h_0+h_\textrm{int})}\approx & {} \bigl (\, e^{-i\,dt\,h_\textrm{int}}\; e^{-i\,dt\,h_0} \,\bigr )^k \end{aligned}$$
(1.17)

where the approximate equality becomes exact in the limit \(dt\rightarrow 0\) which we implicitely assume from now on. The action of \(e^{-i\,dt\,h_0}\) is given by

$$\begin{aligned} (e^{-i\,dt\,h_0}f)(z)= & {} f\bigl ( e^{-i\, dt\, \varepsilon }z\bigr ) \end{aligned}$$
(1.18)

The action of \(e^{-i\,dt\,h_\textrm{int}}\) we write as follows: First, on monomials \(\mathop \Pi _jz_j^{n_j}\) we have

$$\begin{aligned} e^{-i\,dt\,h_\textrm{int}}\,\mathop \Pi _j z_j^{n_j}= & {} e^{iudt\,{\sum }_jn_j}\, e^{-iudt\,{\sum }_jn_j^2}\; \mathop \Pi _j z_j^{n_j} \end{aligned}$$
(1.19)

Recall the Fresnel integral

$$\begin{aligned} \int _{\mathbb R} e^{-i\lambda \phi } \, e^{i{\phi ^2\over 2}} \,{d\phi \over \sqrt{2\pi i}}= & {} e^{-i{\lambda ^2\over 2}} \end{aligned}$$
(1.20)

with \(\,\sqrt{i}=e^{i{\pi \over 4}}\,\). We can write

$$\begin{aligned} e^{-i\,dt\,h_\textrm{int}}\,\mathop \Pi _j z_j^{n_j}= & {} \int _{\mathbb R^{|\Gamma |}} \,\mathop \Pi _j\, \bigl (\, e^{iudt\,-\,i\sqrt{2udt}\, \phi _j} z_j \,\bigr )^{n_j} \, \mathop \Pi _j \, e^{i{\phi _j^2\over 2}} \, {d\phi _j \over \sqrt{2\pi i}} \end{aligned}$$
(1.21)

Thus, for any analytic f (by slight abuse of notation, we temporarily label the lattice sites with natural numbers from 1 to \({|\Gamma |}\) in the middle of the next line)

$$\begin{aligned} f(z)\;=\;f(\,\{z_j\}\,)\!=\! & {} \sum \limits _{n_1,...,n_{|\Gamma |}\!=\!0}^\infty \,c_{n_1\cdots n_{|\Gamma |}}\,z_1^{n_1}\cdots z_{|\Gamma |}^{n_{|\Gamma |}} \;\; \!\equiv \! \;\; \sum \limits _{\{n_j\}} \, c_{\{n_j\}}\; \mathop \Pi _j z_j^{n_j} \end{aligned}$$
(1.22)

we have

$$\begin{aligned} (e^{-i\,dt\,h_\textrm{int}}f)(z)= & {} \sum \limits _{\{n_j\}} \, c_{\{n_j\}}\; e^{-i\,dt\,h_\textrm{int}}\mathop \Pi _j z_j^{n_j} \nonumber \\{} & {} \nonumber \\= & {} \sum \limits _{\{n_j\}} \, c_{\{n_j\}}\; \int _{\mathbb R^{|\Gamma |}} \,\mathop \Pi _j\, \bigl (\, e^{iudt\,-\,i\sqrt{2udt}\, \phi _j} z_j \,\bigr )^{n_j} \, \mathop \Pi _j \, e^{i{\phi _j^2\over 2}} \, {d\phi _j \over \sqrt{2\pi i}} \nonumber \\{} & {} \nonumber \\= & {} \int _{\mathbb R^{|\Gamma |}} f\bigl ( \, \bigl \{\, e^{iudt\,-\,i\sqrt{2udt}\, \phi _j} z_j \,\bigr \}_{j\in \Gamma }\, \bigr ) \; \mathop \Pi _j e^{i{\phi _j^2\over 2}} {d\phi _j \over \sqrt{2\pi i}} \nonumber \\{} & {} \nonumber \\=: & {} \int _{\mathbb R^{|\Gamma |}} f( e^{-i\,dD_{\phi }} z )\;\mathop \Pi _j e^{i{\phi _j^2\over 2}} {d\phi _j \over \sqrt{2\pi i}} \\ \nonumber \end{aligned}$$
(1.23)

with the \({|\Gamma |}\times {|\Gamma |}\) diagonal matrix

$$\begin{aligned} dD_{\phi }:= & {} \textrm{diag}\bigl ( \, \{\, \sqrt{2udt} \,\phi _{j}-udt\, \}_{j\in \Gamma } \,\bigr ) \;\;\in \;\;\mathbb C^{{|\Gamma |}\times {|\Gamma |}} \\ \nonumber \end{aligned}$$
(1.24)

Thus, a single Trotter step is given by

$$\begin{aligned} (e^{-i\,dt\,h}f)(z)= & {} \bigl [ e^{-i\,dt\,h_\textrm{int}}\, (e^{-i\,dt\,h_0}f)\bigr ](z) \nonumber \\{} & {} \nonumber \\= & {} \int _{\mathbb R^{|\Gamma |}} (e^{-i\,dt\,h_0}f)( e^{-i \,dD_{\phi }} z ) \;\mathop \Pi _j e^{i{\phi _j^2\over 2}}{d\phi _j \over \sqrt{2\pi i}} \nonumber \\{} & {} \nonumber \\= & {} \int _{\mathbb R^{|\Gamma |}} f( e^{-i dt \,\varepsilon } e^{-i\,dD_{\phi }} z ) \,\; e^{\,i{\sum }_j{\phi _j^2\over 2}} \mathop \Pi _j {d\phi _j \over \sqrt{2\pi i}} \nonumber \\{} & {} \nonumber \\=: & {} \int _{\mathbb R^{|\Gamma |}} f( e^{-i dt \,\varepsilon } e^{-i\,dD_{\phi }} z ) \,\; e^{\,i{\phi ^2\over 2}} {d^{|\Gamma |}\phi \over (2\pi i)^{{|\Gamma |}/2}} \\ \nonumber \end{aligned}$$
(1.25)

Iterating,

$$\begin{aligned} (e^{-i\,kdt\,h}f)(z)= & {} \int _{\mathbb R^{k{|\Gamma |}}} f\bigl (\, e^{-idt \,\varepsilon } e^{-i\,dD_{\phi _{1}} }\cdots e^{-i dt \,\varepsilon } e^{-i\,dD_{\phi _k} } \,z\, \bigr ) \,\; \mathop \Pi _{\ell =1}^k e^{\,i{\phi _\ell ^2\over 2}} {d^{|\Gamma |}\phi _\ell \over (2\pi i)^{{|\Gamma |}/2}} \\ \nonumber \end{aligned}$$
(1.26)

with the notations

$$\begin{aligned} \phi _\ell:= & {} (\,\{\phi _{j,\ell }\}_{j\in \Gamma }\,)\;\;\in \;\;\mathbb R^{|\Gamma |} \nonumber \\ \phi _\ell ^2:= & {} {\sum }_j\phi _{j,\ell }^2 \;\;\in \;\;\mathbb R \\ d^{|\Gamma |}\phi _\ell:= & {} \mathop \Pi _j d\phi _{j,\ell } \nonumber \end{aligned}$$
(1.27)

That is, on the (real, scalar) integration variable \(\phi _{j,\ell }\in {\mathbb R}\), the first index \(j=(j_1,\cdots ,j_d)\) is a lattice site index and the second index \(\ell \in \mathbb N\) is a time index. Furthermore

$$\begin{aligned} dD_{\phi _\ell }= & {} \textrm{diag}\Bigl ( \; \bigl \{\, \sqrt{2udt} \,\phi _{j,\ell }\,-\,udt\,\bigr \}_{j\in \Gamma } \; \Bigr ) \;\;\in \;\;\mathbb R^{{|\Gamma |}\times {|\Gamma |}} \end{aligned}$$
(1.28)

Let us summarize the above formula in part (a) of the following

Theorem 1

Let \(\;h=h_0+h_\textrm{int}:\mathcal{F}\rightarrow \mathcal{F}\;\) be the Bose-Hubbard Hamiltonian given by (1.12) and (1.13) and let \(\psi \in \mathcal{F}\) be any initial state. Then:

  1. a)

    In the limit \(dt\rightarrow 0\) with \(t=t_k=kdt\,\) fixed, there is the formula

    $$\begin{aligned} (e^{-ikdt\,h}\psi )(z)= & {} \textsf{E}\bigl [ \, \psi ( U_{kdt} z )\,\bigr ] \end{aligned}$$
    (1.29)

    with unitary evolution matrix

    $$\begin{aligned} U_{kdt}:= & {} e^{-i dt \,\varepsilon }\, e^{-i\,dD_{\phi _{1}} }\;\cdots \; e^{-i dt \,\varepsilon }\, e^{-i\,dD_{\phi _k} }\;\; \in \;\; \mathbb {C}^{{|\Gamma |}\times {|\Gamma |}} \end{aligned}$$
    (1.30)

    and Fresnel expectation value

    $$\begin{aligned} \textsf{E}\bigl [ \,\; \cdot \;\,\bigr ]:= & {} \int _{\mathbb R^{k{|\Gamma |}}} \;\cdot \;\; \mathop \Pi _{\ell =1}^k e^{i{\phi _\ell ^2\over 2} } {d^{{|\Gamma |}}\phi _\ell \over (2\pi i)^{{|\Gamma |}/2}} \\ \nonumber \end{aligned}$$
    (1.31)
  2. b)

    The evolution matrix \(U_{t_k}=U_{kdt}\) is a solution of the following stochastic differential equation (SDE)

    $$\begin{aligned} dU_{t_k}\;\;:=\;\;U_{t_k}-U_{t_{k-1}}= & {} -\,i\,U_{t_{k-1}}\,\bigl (\,\varepsilon \, dt\;+\;\sqrt{2u} \;dx_{t_k}\,\bigr ) \end{aligned}$$
    (1.32)

    where \(dx_{t_k}\) is the diagonal matrix of Fresnel Brownian motions given by (1.44) below.

Let’s consider part (b) of the theorem. Suppose we would have Gaussian densities, let’s say 1-dimensional, \(e^{-\phi _{\ell }^{2}/2} \;{d\phi _\ell / \sqrt{2\pi }}\;\) instead of Fresnel kernels \(\; e^{\,i\,{\phi _\ell ^2/ 2}} \;{d\phi _\ell / \sqrt{2\pi i}}\;\). Then the combination of integration variables

$$\begin{aligned} x_{t_k}:= & {} \sqrt{dt}\,{\sum }_{\ell =1}^k \phi _\ell \end{aligned}$$
(1.33)

would be a standard Brownian motion, the product (with \(x_{t_0}:=0\))

$$\begin{aligned} dW:= & {} \mathop \Pi _{\ell =1}^k e^{\,-{\phi _\ell ^2\over 2} }\; { {d\phi _\ell \over \sqrt{2\pi } }} \;\;=\;\; \mathop \Pi _{\ell =1}^k e^{\,-\,{(x_{t_\ell }-x_{t_{\ell -1}})^2\over {2dt}} }\; {{dx_{t_\ell } \over \sqrt{2\pi \,dt} }} \;\;=:\;\; \mathop \Pi _{\ell =1}^k \,p_{dt}(x_{t_\ell },x_{t_{\ell -1}}) \;dx_{t_\ell } \end{aligned}$$
(1.34)

would be standard Wiener measure and in the limit \(dt\rightarrow 0\) there would be the standard Brownian motion calculation rule (Appendix A.1 has a quick reminder)

$$\begin{aligned} (dx_t)^2= & {} dt \end{aligned}$$
(1.35)

and \(\,dx_t\,dt\,=\,dt\,dt\,=\,0\,\). In the presence of Fresnel kernels, we can make the same definitons. That is, the combination of integration variables

$$\begin{aligned} x_{t_k}:= & {} \sqrt{dt}\,{\sum }_{\ell =1}^k \phi _\ell \end{aligned}$$
(1.36)

we call a one-dimensional Fresnel Brownian motion if the \(\phi _\ell \)’s are to be integrated against the product of one-dimensional Fresnel kernels (again with \(x_{t_0}:=0\,\))

$$\begin{aligned} dF:= & {} \mathop \Pi _{\ell =1}^k e^{\,i{\phi _\ell ^2\over 2} } \; {{d\phi _\ell \over \sqrt{2\pi i} }} \;\;=\;\; \mathop \Pi _{\ell =1}^k e^{\,i\,{(x_{t_\ell }-x_{t_{\ell -1}})^2\over {2dt}} }\; {{dx_{t_\ell } \over \sqrt{2\pi i \,dt} }} \;\;=:\;\; \mathop \Pi _{\ell =1}^k \,q_{dt}(x_{t_\ell },x_{t_{\ell -1}}) \;dx_{t_\ell } \end{aligned}$$
(1.37)

which we then refer to as Fresnel measure. Observe that for both Fresnel and Gaussian kernels we have the equations

$$\begin{aligned} \int _{\mathbb R} k_t(x,y)\, k_s(y,z)\, dy= & {} k_{t+s}(x,z) \end{aligned}$$
(1.38)
$$\begin{aligned} \int _{\mathbb R} k_t(x,y)\, dy= & {} 1 \end{aligned}$$
(1.39)

with \(k_t\in \{p_t,q_t\}\). Now, what we can use is the fact that in the limit \(dt\rightarrow 0\) there are analog calculation rules for Fresnel Brownian motions. That is, there are the following formulae (see Appendices A.1 and A.4 for more background)

$$\begin{aligned} (dx_t)^2= & {} i\,dt \end{aligned}$$
(1.40)

and \(\,dx_t\,dt\,=\,dt\,dt\,=\,0\,\).

Now let’s return to the time evolution formula (1.29) of Theorem 1. Instead of 1-dimensional Fresnel kernels we have \({|\Gamma |}\)-dimensional Fresnel kernels and accordingly \({|\Gamma |}\)-dimensional Fresnel Brownian motions

$$\begin{aligned} x_{j,t_k}\;\;=\;\; x_{j,kdt}:= & {} \sqrt{dt}\,{\sum }_{\ell =1}^k \phi _{j,\ell } \end{aligned}$$
(1.41)

where \(j\in \Gamma \) again denotes some lattice site. With

$$\begin{aligned} dx_{j,kdt}:= & {} x_{j,kdt}\,-\,x_{j,(k-1)dt}\;\;=\;\;\sqrt{dt}\,\phi _{j,k} \end{aligned}$$
(1.42)

we can write

$$\begin{aligned} dD_{\phi _k}= & {} \textrm{diag}\bigl ( \; \{\,\sqrt{2u} \,dx_{j,kdt}\,-\,u\,dt\,\}_{j\in \Gamma } \;\bigr ) \nonumber \\=: & {} \sqrt{2u} \;dx_{kdt}\,\;-\,\;u\,dt\,{I\!d} \end{aligned}$$
(1.43)

where we introduced the diagonal matrix of Fresnel Brownian motions

$$\begin{aligned} dx_{kdt}:= & {} \textrm{diag}\bigl (\, \{\, dx_{j,kdt} \,\}_{j\in \Gamma }\,\bigr ) \;\;\in \;\; \mathbb {C}^{{|\Gamma |}\times {|\Gamma |}} \end{aligned}$$
(1.44)

and \(\,{I\!d}\,\) is the \({|\Gamma |}\times {|\Gamma |}\) identity matrix. From the calculation rule (1.40) we get the matrix equation

$$\begin{aligned} (dx_{kdt})^2= & {} \textrm{diag}\bigl (\; \bigl \{\, (dx_{j,kdt})^2 \, \bigr \}_{j\in \Gamma } \; \bigr ) \;\;=\;\; i\,dt\,{I\!d} \end{aligned}$$
(1.45)

Thus, up to terms \(\;O\bigl (\,dt^{3/2}\,\bigr )\,\),

$$\begin{aligned} e^{-i\,dD_{\phi _k}}= & {} 1\;-\;i\,dD_{\phi _k}\;-\; {1\over 2}\,(dD_{\phi _k})^2 \nonumber \\= & {} 1\;-\;i\sqrt{2u} \,dx_{kdt}\;+\;i\,u\,dt\,{I\!d}\;-\; {1\over 2}\,\bigl ( \sqrt{2u} \,dx_{kdt}\;-\;u\,dt\,{I\!d}\bigr )^2 \nonumber \\= & {} 1\;-\;i\sqrt{2u} \,dx_{kdt}\;+\;i\,u\,dt\,{I\!d}\;-\; {1\over 2}\,\bigl ( \sqrt{2u} \,dx_{kdt} \bigr )^2 \nonumber \\= & {} 1\;-\;i\sqrt{2u} \,dx_{kdt} \end{aligned}$$
(1.46)

and we arrive at the following SDE for the evolution matrix \(U_t\):

$$\begin{aligned} U_{kdt}= & {} e^{-i dt \,\varepsilon } e^{-i\,dD_{\phi _{1}} }\;\times \;\cdots \;\times \;e^{-i dt \,\varepsilon } e^{-i\,dD_{\phi _{k-1}} }\;\times \; e^{-i dt \,\varepsilon } e^{-i\,dD_{\phi _k} } \nonumber \\= & {} U_{(k-1)dt} \;\times \; e^{-i dt \,\varepsilon } e^{-i\,dD_{\phi _k} } \nonumber \\&\buildrel \text {(1.46)}\over =&U_{(k-1)dt} \;\bigl (\, 1\;-\;i dt\,\varepsilon \,\bigr )\;\bigl (\,1\;-\;i\sqrt{2u} \;dx_{kdt}\,\bigr )\; \nonumber \\= & {} U_{(k-1)dt} \;\bigl (\, 1\;-\;i dt\,\varepsilon \;-\;i\sqrt{2u} \;dx_{kdt}\,\bigr )\; \end{aligned}$$
(1.47)

or, with \(t_k=kdt\),

$$\begin{aligned} dU_{t_k}\;\;:=\;\;U_{t_k}-U_{t_{k-1}}= & {} -i\,U_{t_{k-1}}\,\bigl (\, dt\,\varepsilon \;+\;\sqrt{2u} \;dx_{t_k}\,\bigr ) \end{aligned}$$
(1.48)

which completes the derivation of part (b) of Theorem 1. More compactly, this could be written as

$$\begin{aligned} dU_t= & {} -i\,U_t\,\bigl (\,\varepsilon \, dt\;+\;\sqrt{2u} \;dx_t\,\bigr ) \end{aligned}$$
(1.49)

but we want to remind at this place that when discretizing stochastic differential equations or stochastic integrals it is usually crucial whether a particular time index is a \(t_k\) or a \(t_{k-1}\). Throughout this paper, we use the Ito definition which is as follows: If some quantity \(Q_t\) satisfies the stochastic differential equation

$$\begin{aligned} dQ_t= & {} A_t\,dt \;+\;B_t\,dx_t \end{aligned}$$
(1.50)

then this is equivalent to the following discrete time update rule

$$\begin{aligned} { Q_{t_k}(\phi _1,\cdots ,\phi _k)\;\;=\;\;} \\{} & {} Q_{t_{k-1}}(\phi _1,\cdots ,\phi _{k-1}) \;+\;A_{t_{k-1}}(\phi _1,\cdots ,\phi _{k-1})\,dt \;+\;B_{t_{k-1}}(\phi _1,\cdots ,\phi _{k-1})\,\sqrt{dt}\;\phi _k \nonumber \end{aligned}$$
(1.51)

That is, the new random number or integration variable \(\phi _k\) which enters when going from time \(t_{k-1}\) to \(t_k\) enters in an explicitely given way, namely through the explicit \(\phi _k\) on the very right of (1.51). There are no \(\phi _k\)’s in the A, B or \(Q_{t_{k-1}}\) on the right hand side of (1.51). This then for example has the immediate consequence that

$$\begin{aligned} \textsf{E}[\,Q_{t_k}\,]= & {} \textsf{E}[\,Q_{t_{k-1}}\,]\;+\;\textsf{E}[\,A_{t_{k-1}}\,]\,dt \end{aligned}$$
(1.52)

since the diffusive part does not contribute to the expectation value because of \(\,\textsf{E}[\phi _k]=0\,\). Thus, quantities \(Q_t\) which have a vanishing drift part \(A=0\) are of special importance since their expectation value does not change over time and they are called martingales. Appendix A.2 has a quick reminder on Ito and Stratonovich integrals and why there are different definitions at all.

1.2 Density Matrix Elements from Stochastic Differential Equations

We consider the following normalized initial state with \({\lambda }=\{{\lambda }_j\}\in \mathbb C^{|\Gamma |}\)

$$\begin{aligned} \psi _0(z)\;=\;\psi _0(\,\{z_j\}\,):= & {} \prod _j \,e^{{\lambda }_j z_j}\; e^{-{|{\lambda }_j|^2\over 2}} \;\;=:\;\; e^{{\lambda }z} \;e^{-{|{\lambda }|^2\over 2}} \end{aligned}$$
(1.53)

This is a product of coherent states. The expected number of particles at site j is

$$\begin{aligned} (\,\psi _0\,,a_j^+a_j \psi _0\,)_\mathcal{F}= & {} |{\lambda }_j|^2 \end{aligned}$$
(1.54)

which means that \(\;|{\lambda }_j|=\sqrt{N_j}\;\) would be a natural choice if we want to have \(N_j\) particles at site j. The total number of particles N is given by

$$\begin{aligned} N\;\;=\;\; {\sum }_j N_j= & {} {\sum }_j |\lambda _j|^2\;\;=\;\;|\lambda |^2 \end{aligned}$$
(1.55)

Let’s consider the time evolution of this \(\psi _0\). According to Theorem 1, we have

$$\begin{aligned} \psi _t(z)\;=\;(e^{-ith}\psi _0)(z)= & {} \textsf{E}[\,\psi _0(U_tz)\,] \end{aligned}$$
(1.56)

where the evolution matrix \(U_t\) is given by the SDE

$$\begin{aligned} dU_t= & {} -\,i\,U_t\,\bigl (\, dt\,\varepsilon \;+\;\sqrt{2u} \;dx_t\,\bigr ) \end{aligned}$$
(1.57)

with initial condition \(\,U_0={I\!d}\,\) and \(x_t\) being a Fresnel Brownian motion. We want to calculate the time evolution of the density matrix elements

$$\begin{aligned} (\,\psi _t\,,a_i^+a_j \,\psi _t\,)_\mathcal{F}= & {} (\,\psi _t\,,[a_j a_i^+ -\delta _{i,j}] \,\psi _t\,)_\mathcal{F}\nonumber \\= & {} (\,a_j^+\psi _t\,, a^+_i \psi _t\,)_\mathcal{F}\;-\; \delta _{i,j}\,(\,\psi _t\,, \psi _t\,)_\mathcal{F}\nonumber \\= & {} \int _{\mathbb {C}^{|\Gamma |}} z_j {\bar{z}}_i\; |\psi _t(z)|^2 \,d\mu (z) \;\;-\;\;\delta _{i,j} \end{aligned}$$
(1.58)

From Theorem 1 we have the representations (with  \(t=t_k=kdt\) )

$$\begin{aligned} \psi _t(z)= & {} \textsf{E}[\,\psi _0(U_tz)\,]\;\;=\;\; \int \,\psi _0(U_{x,t}z)\; dF(\{x_t\}) \\ \overline{\psi _t(z)}= & {} \overline{\textsf{E}[\,\psi _0(U_tz)\,]}\;\;=\;\; \int \,\overline{\psi _0(U_{y,t}z)}\; d\bar{F}(\{y_t\}) \nonumber \end{aligned}$$
(1.59)

with

$$\begin{aligned} dF= & {} \mathop \Pi _{\ell =1}^k e^{\,+\,i{\phi _\ell ^2\over 2}} { {d^{|\Gamma |}\phi _\ell \over (2\pi i)^{{|\Gamma |}/2}} } \;\;=\;\; \mathop \Pi _{\ell =1}^k e^{\,+\,i{(x_{t_\ell }-x_{t_{\ell -1}})^2\over 2 dt}} {d^{|\Gamma |} x_{t_\ell } \over (2\pi i\,dt)^{{|\Gamma |}/2}} \\ d\bar{F}= & {} \mathop \Pi _{\ell =1}^k e^{\,-\,i{\theta _\ell ^2\over 2}} { {d^{|\Gamma |}\theta _\ell \over [2\pi (-i)]^{{|\Gamma |}/2}} } \;\;=\;\; \mathop \Pi _{\ell =1}^k e^{\,-\,i{(y_{t_\ell }-y_{t_{\ell -1}})^2\over 2 dt}} {d^{|\Gamma |} y_{t_\ell } \over [2\pi (-i)dt]^{{|\Gamma |}/2}} \nonumber \end{aligned}$$
(1.60)

where \(\,\sqrt{-i}\,:=\,\overline{\sqrt{i}}\,=\,e^{\,-\,i{\pi \over 4}}\,\) and \(x_{t_\ell }\) and \(y_{t_\ell }\) are \({|\Gamma |}\)-dimensional Fresnel Brownian motions given by

$$\begin{aligned} x_{t_\ell }= & {} \sqrt{dt}\,{\sum }_{m=1}^\ell \phi _m \\ y_{t_\ell }= & {} \sqrt{dt}\,{\sum }_{m=1}^\ell \theta _m \nonumber \end{aligned}$$
(1.61)

With that, we can write

$$\begin{aligned} { \int _{\mathbb C^{|\Gamma |}} z_j {\bar{z}}_i\; |\psi _t(z)|^2 \,d\mu (z)} \\= & {} \int _{\mathbb C^{|\Gamma |}} z_j {\bar{z}}_i\; \int \,\psi _0(U_{x,t}z)\; dF(\{x_t\})\; \int \,\overline{\psi _0(U_{y,t}z)}\; d\bar{F}(\{y_t\}) \;d\mu (z) \nonumber \\= & {} \int \int \;\Bigl \{\; \int _{\mathbb C^{|\Gamma |}} z_j {\bar{z}}_i\; \psi _0(U_{x,t}z)\;\overline{\psi _0(U_{y,t}z)}\;d\mu (z)\;\Bigr \} \; dF(\{x_t\})\; d\bar{F}(\{y_t\}) \nonumber \end{aligned}$$
(1.62)

and in the same way

$$\begin{aligned} \Vert \psi _t\Vert _\mathcal{F}^2= & {} \int _{\mathbb C^{|\Gamma |}} |\psi _t(z)|^2 \,d\mu (z) \\= & {} \int \int \;\Bigl \{\; \int _{\mathbb C^{|\Gamma |}} \psi _0(U_{x,t}z)\;\overline{\psi _0(U_{y,t}z)}\;d\mu (z)\;\Bigr \} \; dF(\{x_t\})\; d\bar{F}(\{y_t\}) \nonumber \end{aligned}$$
(1.63)

The wavy brackets above are the expectations over the bosonic Fock space, written in the Bargmann-Segal representation, and can be calculated. Since

$$\begin{aligned} \psi _0(U_{x,t}z)\;\overline{\psi _0(U_{y,t}z)}= & {} \exp \bigl \{\, U_{x,t}^T{\lambda }\cdot z \,\bigr \}\;\exp \bigl \{\, {\bar{U}}_{y,t}^T{\bar{\lambda }}\cdot {\bar{z}}\,\bigr \} \; e^{-|{\lambda }|^2} \end{aligned}$$
(1.64)

with \(U^T\) denoting the transpose of U, and because of the formulae

$$\begin{aligned} \int _{\mathbb C^{|\Gamma |}} e^{\lambda z+\bar{\lambda }\bar{z}}\; d\mu (z)= & {} e^{\lambda \bar{\lambda }} \nonumber \\ \int _{\mathbb C^{|\Gamma |}} z_j\,{\bar{z}}_i\; e^{\lambda z+\bar{\lambda }\bar{z}}\;d\mu (z)= & {} {\partial \over \partial \lambda _j} {\partial \over \partial {\bar{\lambda }_i}}\;e^{\lambda \bar{\lambda }} \;\; = \;\; ({\lambda }_i\,{\bar{\lambda }}_j+\delta _{i,j})\;e^{\lambda \bar{\lambda }} \end{aligned}$$
(1.65)

we obtain the following representations: The norm of \(\psi _t\) is given by

$$\begin{aligned} \Vert \psi _t\Vert _\mathcal{F}^2= & {} \textsf{E}_x\bar{\textsf{E}}_y\Bigl [ \; \exp \bigl \{\, U_{x,t}^T{\lambda }\cdot {\bar{U}}_{y,t}^T{\bar{\lambda }}\,\bigr \} \;\Bigr ] \; e^{-|{\lambda }|^2} \end{aligned}$$
(1.66)

and density matrix elements can be written as

$$\begin{aligned} (\,\psi _t\,,a_i^+a_j \,\psi _t\,)_\mathcal{F}= & {} \textsf{E}_x\bar{\textsf{E}}_y\Bigl [\, \;[U_{x,t}^T{\lambda }]_i\;\, [{\bar{U}}_{y,t}^T{\bar{\lambda }}]_j\; \exp \bigl \{\, U_{x,t}^T{\lambda }\cdot {\bar{U}}_{y,t}^T{\bar{\lambda }}\,\bigr \} \;\Bigr ] \;e^{-|{\lambda }|^2} \end{aligned}$$
(1.67)

where we used the notation

$$\begin{aligned} \textsf{E}_x\bar{\textsf{E}}_y\bigl [\;\cdots \;\bigr ]= & {} \int \int \;\cdots \; dF(\{x_t\})\; d\bar{F}(\{y_t\}) \end{aligned}$$
(1.68)

for the Fresnel expectations.

Now recall that

$$\begin{aligned} dU_{x,t}= & {} -\,i\,U_{x,t}\,\bigl (\, dt\,\varepsilon \;+\;\sqrt{2u} \,dx_{t}\,\bigr ) \end{aligned}$$
(1.69)

or, since \(\varepsilon _{ij}=\varepsilon _{ji}\,\) or \(\,\varepsilon ^T=\varepsilon \),

$$\begin{aligned} dU_{x,t}^T= & {} -\,i\,\bigl (\, dt\,\varepsilon \;+\;\sqrt{2u} \,dx_{t}\,\bigr )\,U_{x,t}^T \end{aligned}$$
(1.70)

Thus,

$$\begin{aligned} d(U_{x,t}^T{\lambda })= & {} -\,i\,\bigl (\, dt\,\varepsilon \;+\;\sqrt{2u} \,dx_{t}\,\bigr )\,U_{x,t}^T{\lambda } \end{aligned}$$
(1.71)

and in the same way

$$\begin{aligned} d({\bar{U}}_{y,t}^T{\bar{\lambda }})= & {} +\,i\,\bigl (\, dt\,\varepsilon \;+\;\sqrt{2u} \,dy_{t}\,\bigr )\,{\bar{U}}_{y,t}^T{\bar{\lambda }} \end{aligned}$$
(1.72)

Thus, if we abbreviate the quantities (where the \(\bar{v}\) at this stage is not the complex conjugate of v since it has different integration variables)

$$\begin{aligned} v\;=\;v_{x,t}:= & {} U_{x,t}^T{\lambda }\;\;\in \;\;\mathbb C^{|\Gamma |} \\ {\bar{v}}\;=\;{\bar{v}}_{y,t}:= & {} {\bar{U}}_{y,t}^T{\bar{\lambda }}\;\;\in \;\;\mathbb C^{|\Gamma |} \nonumber \end{aligned}$$
(1.73)

we obtain the SDE system

$$\begin{aligned} dv= & {} \;-\;i\,\bigl (\, dt\,\varepsilon \;+\;\sqrt{2u} \,dx_{t}\,\bigr )\,v \\ d{\bar{v}}= & {} \; +\; i\,\bigl (\, dt\,\varepsilon \;+\;\sqrt{2u} \,dy_{t}\,\bigr )\,{\bar{v}}\nonumber \end{aligned}$$
(1.74)

with initial conditions

$$\begin{aligned} v_0= & {} {\lambda }\\ {\bar{v}}_0= & {} {\bar{\lambda }}\nonumber \end{aligned}$$
(1.75)

In coordinates, this reads

$$\begin{aligned} dv_j= & {} \;-\;i\, dt\,(\varepsilon v)_j \;-\;i\,\sqrt{2u}\,v_j\,dx_j \\ d\bar{v}_j= & {} \;+\;i\, dt\,(\varepsilon \bar{v})_j \;+\;i\,\sqrt{2u}\,\bar{v}_j\,dy_j \nonumber \end{aligned}$$
(1.76)

or even more explicitely, making also the times and the Fresnel integration variables x and y explicit at the \(v,{\bar{v}}\),

$$\begin{aligned} dv_{j,x,t}= & {} \;-\;i\, dt\,{\sum }_i\varepsilon _{ji}\, v_{i,x,t} \;-\;i\,\sqrt{2u}\,v_{j,x,t}\,dx_{j,t} \\ d\bar{v}_{j,y,t}= & {} \;+\;i\, dt\,{\sum }_i\varepsilon _{ji}\,\bar{v}_{i,y,t} \;+\;i\,\sqrt{2u}\,\bar{v}_{j,y,t}\,dy_{j,t} \nonumber \end{aligned}$$
(1.77)

Then, from (1.67) we obtain the density matrix elements as

$$\begin{aligned} (\,\psi _t\,,a_i^+a_j \,\psi _t\,)_\mathcal{F}= & {} \textsf{E}_x\bar{\textsf{E}}_y\bigl [ \;v_i \,{\bar{v}}_j\; e^{v{\bar{v}}} \;\bigr ] \;e^{-|{\lambda }|^2} \end{aligned}$$
(1.78)

if we use the notation

$$\begin{aligned} v{\bar{v}}:= & {} {\sum }_j\,v_j\bar{v}_j \;\;=\;\; v^T{\bar{v}} \end{aligned}$$
(1.79)

Finally, the norm of \(\psi _t\) has the representation

$$\begin{aligned} \Vert \psi _t\Vert ^2_\mathcal{F}= & {} \textsf{E}_x\bar{\textsf{E}}_y[ \; e^{v{\bar{v}}} \;] \;e^{-|{\lambda }|^2} \end{aligned}$$
(1.80)

We summarize the results in the following

Theorem 2

Let \(\psi _0\) be the initial state

$$\begin{aligned} \psi _0(z)\;=\;\psi _0(\,\{z_j\}\,)= & {} \prod _j \,e^{{\lambda }_j z_j}\; e^{-{|{\lambda }_j|^2\over 2}} \;\;=:\;\; e^{{\lambda }z} \;e^{-{|{\lambda }|^2\over 2}} \end{aligned}$$
(1.81)

and let \(\,\psi _t=e^{-ith}\psi _0\,\) be the time evolved state with Bose-Hubbard Hamiltonian \(h=h_0+h_\textrm{int}\) given by (1.12) and (1.13). Then there are the following representations:

$$\begin{aligned} (\,\psi _t\,,a_i^+a_j \,\psi _t\,)_\mathcal{F}= & {} \textsf{E}_x\bar{\textsf{E}}_y\bigl [ \;v_{i,x,t} \;{\bar{v}}_{j,y,t}\; e^{v_{x,t}{\bar{v}}_{y,t}} \;\bigr ] \;e^{-|{\lambda }|^2} \end{aligned}$$
(1.82)
$$\begin{aligned} \Vert \psi _t\Vert ^2_\mathcal{F}= & {} \textsf{E}_x\bar{\textsf{E}}_y[ \; e^{v_{x,t}{\bar{v}}_{y,t}} \;] \;e^{-|{\lambda }|^2} \end{aligned}$$
(1.83)

with Fresnel expectations \(\,\textsf{E}_x\bar{\textsf{E}}_y[ \;\cdot \;]\,\) given by (1.68) and (1.60) above, and the \(v,{\bar{v}}\in \mathbb C^{|\Gamma |}\,\) are given by the SDE system

$$\begin{aligned} dv_{j}= & {} \;-\;i\, dt\,(\varepsilon v)_j \;-\;i\,\sqrt{2u}\,v_j\,dx_j \nonumber \\ d\bar{v}_{j}= & {} \;+\;i\, dt\,(\varepsilon \bar{v})_j \;+\;i\,\sqrt{2u}\,\bar{v}_j\,dy_j \end{aligned}$$
(1.84)

with initial conditions \(\,v_{x,0}={\lambda },\,{\bar{v}}_{y,0}=\bar{\lambda }\,\).

Now, in the next chapter we will see that the combination \(v_{x,t}{\bar{v}}_{y,t}\) is a martingale and as a consequence, the exponential \(\,e^{v_{x,t}{\bar{v}}_{y,t}}\,\) can be absorbed into the Fresnel integration measure. This has the effect that the density matrix elements are then simply given by

$$\begin{aligned}(\,\psi _t\,,a_i^+a_j \,\psi _t\,)_\mathcal{F}= & {} \textsf{E}_{\tilde{x}}\bar{\textsf{E}}_{\tilde{y}}\bigl [ \;v_{i,\tilde{x},t} \;{\bar{v}}_{j,\tilde{y},t}\;\bigr ] \end{aligned}$$

where the \(\;v_{\tilde{x},t},\,{\bar{v}}_{\tilde{y},t}\;\) are given by the transformed SDE system

$$\begin{aligned}dv_j= & {} \;-\;i\, dt\,(\varepsilon v)_j \;-\;i\,2u\,dt\,v_j{\bar{v}}_j\,v_j \;-\;i\,\sqrt{2u}\,v_j\,d\tilde{x}_j \\ d\bar{v}_j= & {} \;+\;i\, dt\,(\varepsilon \bar{v})_j\;+\;i\,2u\,dt\,v_j{\bar{v}}_j\,{\bar{v}}_j \;+\;i\,\sqrt{2u}\,\bar{v}_j\,d\tilde{y}_j \end{aligned}$$

with transformed Fresnel Brownian motions

$$\begin{aligned}d\tilde{x}_{j,t_\ell }:= & {} dx_{j,t_\ell } \,-\, \sqrt{2u}\,dt\, (v_j{\bar{v}}_j)_{t_{\ell -1}} \\ d\tilde{y}_{j,t_\ell }:= & {} dy_{j,t_\ell } \,-\, \sqrt{2u}\,dt\, (v_j{\bar{v}}_j)_{t_{\ell -1}} \end{aligned}$$

So, let’s look at the details.

2 Martingale Property and Girsanov Transformation

2.1 Unitary Time Evolution as a Martingale

Recall from Theorem 1 that the time evolution \(\,\psi _t=e^{-it(h_0+h_\textrm{int})}\psi _0\,\) of some initial state \(\psi _0=\psi _0(z)\) can be written as a Fresnel expectation value

$$\begin{aligned} \psi _t(z)= & {} \textsf{E}\bigl [\, \psi _0(U_tz) \,\bigr ]\;\;=\;\;\textsf{E}_x\bigl [\, \psi _0(U_{x,t}z) \,\bigr ]\;\;=\;\;\int \psi _0(U_{x,t}z) \,dF(\{x_s\}_{0<s\le t}) \end{aligned}$$
(2.1)

where the unitary evolution matrix is given by the SDE

$$\begin{aligned} dU_{x,t}= & {} -i\,U_{x,t}\,\bigl (\, dt\,\varepsilon \;+\; \sqrt{2u}\, dx_t\,\bigr ) \end{aligned}$$
(2.2)

with initial value \(U_{x,0}={I\!d}\). The norm of \(\psi _t\) is given by

$$\begin{aligned} \Vert \psi _t\Vert _\mathcal{F}^2\;\;=\;\; (\psi _t,\psi _t)_\mathcal{F}= & {} \int _{\mathbb C^{|\Gamma |}} \,\psi _t(z)\, \overline{\psi _t(z)}\, d\mu (z) \end{aligned}$$
(2.3)

with \(d\mu (z)\) given by (1.7). For the complex conjugated \(\psi _t\) we use integration variables or Fresnel Brownian motions \(\{y_s\}_{0<s\le t}\) and write

$$\begin{aligned} \overline{\psi _t(z)}= & {} \bar{\textsf{E}}_y\bigl [\, \overline{\psi _0(U_{y,t}z)} \,\bigr ]\;\;=\;\;\int \,\overline{\psi _0(U_{y,t}z)} \,d\bar{F}(\{y_s\}_{0<s\le t}) \end{aligned}$$
(2.4)

such that, as in (1.63) of the last chapter,

$$\begin{aligned} \Vert \psi _t\Vert _\mathcal{F}^2= & {} \int _{\mathbb C^{|\Gamma |}} \,\psi _t(z)\, \overline{\psi _t(z)}\, d\mu (z) \nonumber \\= & {} \int _{\mathbb C^{|\Gamma |}} \,\textsf{E}_x\bigl [\, \psi _0(U_{x,t}z) \,\bigr ]\; \bar{\textsf{E}}_y\bigl [\, \overline{\psi _0(U_{y,t}z)} \,\bigr ] \; d\mu (z) \nonumber \\= & {} \textsf{E}_x\bar{\textsf{E}}_y\Bigl [\; \int _{\mathbb C^{|\Gamma |}} \, \psi _0(U_{x,t}z) \, \overline{\psi _0(U_{y,t}z)} \; d\mu (z) \;\Bigr ] \end{aligned}$$
(2.5)

Now we have

$$\begin{aligned} d\mu (Uz)= & {} d\mu (z) \end{aligned}$$
(2.6)

for any unitary U. Thus, with the substitution \(z=U_{y,t}^+\,w\,\) with \(U^+=\bar{U}^T\) being the adjoint matrix, and renaming the w back to z, we obtain

$$\begin{aligned} \Vert \psi _t\Vert _\mathcal{F}^2= & {} \textsf{E}_x\bar{\textsf{E}}_y\Bigl [\; \int _{\mathbb C^{|\Gamma |}} \, \psi _0(U_{x,t}U_{y,t}^+z) \, \overline{\psi _0(z)} \; d\mu (z) \;\Bigr ] \end{aligned}$$
(2.7)

This quantity has to be independent of t, we have to have

$$\begin{aligned} \Vert \psi _t\Vert _\mathcal{F}^2= & {} \textsf{E}_x\bar{\textsf{E}}_y\Bigl [\; \int _{\mathbb C^{|\Gamma |}} \, \psi _0(U_{x,t}U_{y,t}^+z) \, \overline{\psi _0(z)} \; d\mu (z) \;\Bigr ] \nonumber \\&\buildrel ! \over =&\int _{\mathbb C^{|\Gamma |}} \, \psi _0(z) \, \overline{\psi _0(z)} \; d\mu (z) \;\;=\;\; \Vert \psi _0\Vert _\mathcal{F}^2 \end{aligned}$$
(2.8)

How can this be understood from a stochastic calculus point of view? We have

$$\begin{aligned} dU_{x,t}= & {} -\,i\,U_{x,t}\,\bigl (\, dt\,\varepsilon \;+\; \sqrt{2u}\, dx_t\,\bigr ) \nonumber \\ dU^+_{y,t}= & {} +\,i\,\bigl (\, dt\,\varepsilon \;+\; \sqrt{2u}\, dy_t\,\bigr )\, U_{y,t}^+ \end{aligned}$$
(2.9)

Since \(\,dx_t\,dy_t=0\,\), we also have \(\,dU_{x,t}\,dU_{y,t}^+=0\,\) such that

$$\begin{aligned} d\bigl (\, U_{x,t}U_{y,t}^+\,\bigr )= & {} dU_{x,t}\,U_{y,t}^+ \;+\; U_{x,t}\, dU_{y,t}^+ \;+\; dU_{x,t}\,dU_{y,t}^+ \nonumber \\= & {} -\,i\,U_{x,t}\,\bigl (\, dt\,\varepsilon \,+\, \sqrt{2u}\, dx_t\,\bigr )\,U_{y,t}^+ \;+\;i\,U_{x,t}\,\bigl (\, dt\,\varepsilon \,+\, \sqrt{2u}\, dy_t\,\bigr )\, U_{y,t}^+ \;\;+\;0 \nonumber \\= & {} \,-\,i\,\sqrt{2u}\; U_{x,t}\,(dx_t-dy_t)\,U_{y,t}^+ \nonumber \\= & {} \,-\,i\,\sqrt{4u}\; U_{x,t}\,d\xi _t\,U_{y,t}^+ \end{aligned}$$
(2.10)

if we put (the eta’s will be used later, not now)

$$\begin{aligned} d\xi _t\;\;:=\;\; {dx_t-dy_t\over \sqrt{2} }\;,\;\;\;\;\;\; d\eta _t\;\;:=\;\; {dx_t+dy_t\over \sqrt{2} } \end{aligned}$$
(2.11)

Equation (2.10) means that the matrix \(\,U_{x,t}U_{y,t}^+\,\) is a martingale, its \(\,d(U_{x,t}U_{y,t}^+)\,\) has no drift part, no dt part, but only a diffusive part, a \(dx_t\) or \(dy_t\) part. And since

$$\begin{aligned} \textsf{E}[\,dx_t\,]\;=\;\bar{\textsf{E}}[\,dy_t\,]\;=\;0 \end{aligned}$$
(2.12)

one then has

$$\begin{aligned} \textsf{E}\bar{\textsf{E}}\bigl [\,d(U_{x,t}U_{y,t}^+)\,\bigr ]= & {} 0 \end{aligned}$$
(2.13)

which results in

$$\begin{aligned} \textsf{E}\bar{\textsf{E}}\bigl [\, U_{x,t}U_{y,t}^+\,\bigr ]= & {} U_{x,0}\,U_{y,0}^+\;\;=\;\;{I\!d} \end{aligned}$$
(2.14)

However, in (2.8) we have not directly an expectation of the matrix \(\,U_{x,t}U_{y,t}^+\,\) itself, but we have an arbitrary function of it, so we have to consider an expectation of the form

$$\begin{aligned} \textsf{E}_x\bar{\textsf{E}}_y\Bigl [\; f\bigl ( \,U_{x,t}U_{y,t}^+ \,\bigr ) \;\Bigr ] \end{aligned}$$
(2.15)

where \(\,f:\mathbb C^{{|\Gamma |}\times {|\Gamma |}}\rightarrow \mathbb C\,\) is an arbitrary function. These quantities should also be time independent, how can we understand that? Let’s abbreviate for the moment

$$\begin{aligned} M_t:= & {} U_{x,t}U_{y,t}^+ \;\;=\;\; (\,M_{ij}\,)_{i,j\in \Gamma } \end{aligned}$$
(2.16)

Then, with the Ito formula (Appendix A.2 has a quick reminder),

$$\begin{aligned} df(M_t)\!=\! & {} {\sum }_{i,j\in \Gamma }\, {\partial f \over \partial M_{ij}}\, dM_{ij} \;\!+\!\; {1\over 2} {\sum }_{i,j\in \Gamma } {\sum }_{k,\ell \in \Gamma } \,{\partial ^2 f \over \partial M_{ij} \partial M_{k\ell }}\, dM_{ij}\, dM_{k\ell } \end{aligned}$$
(2.17)

In the equation above, also the \(\,k,\ell \) are temporarily used as lattice site indices, they are no time indices here. The first sum is purely diffusive since

$$\begin{aligned} dM_{ij}= & {} \,-\,i\,\sqrt{4u}\, [\,U_{x,t}\,d\xi _t\,U_{y,t}^+\,]_{i,j} \end{aligned}$$
(2.18)

has no dt part. And the second sum actually vanishes since

$$\begin{aligned} dM_{ij}\, dM_{k\ell }= & {} \,-\,4u\, [\,U_{x,t}\,d\xi _t\,U_{y,t}^+\,]_{i,j}\; [\,U_{x,t}\,d\xi _t\,U_{y,t}^+\,]_{k,\ell } \nonumber \\= & {} \,-\,4u\, \sum \limits _{m,n\in \Gamma } [U_{x,t}]_{i,m}\,d\xi _{m,t}\,[U_{y,t}^+]_{m,j}\; [U_{x,t}]_{k,n}\,d\xi _{n,t}\,[U_{y,t}^+]_{n,\ell } \nonumber \\&\buildrel d\xi _{m,t}\,d\xi _{n,t}\,=\,0 \over =&\;\;0 \end{aligned}$$
(2.19)

Namely,

$$\begin{aligned} d\xi _{m,t}\,d\xi _{n,t}= & {} {1\over 2}\, (dx_{m,t}-dy_{m,t})\,(dx_{n,t}-dy_{n,t}) \nonumber \\= & {} {1\over 2}\, \bigl (\, dx_{m,t}\,dx_{n,t} \, + \, dy_{m,t}\,dy_{n,t} \, - \, dx_{m,t}\,dy_{n,t} \, -\, dy_{m,t}\,dx_{n,t} \, \bigr ) \nonumber \\= & {} {1\over 2}\, \bigl (\, dx_{m,t}\,dx_{n,t} \, + \, dy_{m,t}\,dy_{n,t} \, - \, 0 \, -\, 0 \, \bigr ) \end{aligned}$$
(2.20)

For \(m=n\), this becomes

$$\begin{aligned} d\xi _{m,t}\,d\xi _{m,t}= & {} {1\over 2}\, \bigl (\, (dx_{m,t})^2 \, + \, (dy_{m,t})^2 \, \bigr ) \nonumber \\= & {} {1\over 2}\, \bigl (\, i\,dt \, - \, i\, dt \, \bigr ) \;\;=\;\; 0 \end{aligned}$$
(2.21)

And for \(m\ne n\), this is simply

$$\begin{aligned} d\xi _{m,t}\,d\xi _{n,t}= & {} {1\over 2}\, \bigl (\, 0 \, + \, 0 \, \bigr )\;\;=\;\; 0 \end{aligned}$$
(2.22)

Thus we end up with

$$\begin{aligned} df(M_t)= & {} \,-\,i\,\sqrt{4u}\,{\sum }_{i,j}\, {\partial f \over \partial M_{ij}}\, [\,U_{x,t}\,d\xi _t\,U_{y,t}^+\,]_{i,j} \\ \nonumber \end{aligned}$$
(2.23)

which is purely diffusive and this results in

$$\begin{aligned} \textsf{E}_x\bar{\textsf{E}}_y\bigl [\,df(M_t)\,\bigr ]= & {} 0 \end{aligned}$$
(2.24)

and accordingly

$$\begin{aligned} \textsf{E}_x\bar{\textsf{E}}_y\bigl [\,f(M_t)\,\bigr ]= & {} f(M_0) \;+\; \int _0^t \, \textsf{E}_x\bar{\textsf{E}}_y\bigl [\,df(M_s)\,\bigr ] \;\;=\;\; f(M_0) \end{aligned}$$
(2.25)

for arbitrary f. Let’s summarize these observations in the following

Theorem 3

Let \(U_{x,t}\) be the unitary evolution matrix of Theorem 1 such that the time evolution of an arbitrary state \(\psi \in \mathcal{F}\) can be written as

$$\begin{aligned} (e^{-ikdt\,h}\psi )(z)= & {} \textsf{E}_x\bigl [ \, \psi ( U_{x,kdt} z )\,\bigr ] \end{aligned}$$
(2.26)

with Fresnel expectation \(\textsf{E}_x[\,\cdot \,]\) given by (1.60) and (1.68). Then, for arbitrary \(\,f:\mathbb C^{{|\Gamma |}\times {|\Gamma |}}\rightarrow \mathbb C\,\), the quantity \(f( U_{x,t}U_{y,t}^+)\) is a martingale, its \(df( U_{x,t}U_{y,t}^+)\) has no dt-part, and we have the following identity:

$$\begin{aligned} \textsf{E}_x\bar{\textsf{E}}_y\Bigl [\; f\bigl ( \,U_{x,t}U_{y,t}^+ \,\bigr ) \;\Bigr ]= & {} f\bigl ( \,U_{0}U_{0}^+ \,\bigr )\;\;=\;\;f\bigl ( \,{I\!d}\,\bigr ) \end{aligned}$$
(2.27)

In particular, for any time evolved state \(\psi _t=e^{-it\,h}\psi _0\,\),

$$\begin{aligned} \Vert \psi _t\Vert _\mathcal{F}^2= & {} \int _{\mathbb C^{|\Gamma |}} \, \textsf{E}_x\bar{\textsf{E}}_y\bigl [\,\psi _0(U_{x,t}U_{y,t}^+z) \, \overline{\psi _0(z)}\,\bigr ] \; d\mu (z) \nonumber \\= & {} \int _{\mathbb C^{|\Gamma |}} \, \psi _0(z) \, \overline{\psi _0(z)} \; d\mu (z) \;\;=\;\; \Vert \psi _0\Vert _\mathcal{F}^2 \;\;. \end{aligned}$$
(2.28)

2.2 Girsanov Transformed SDE System for the Density Matrix Elements

Recall the representations of Theorem 2, for the initial state \(\,\psi _0(z)=e^{{\lambda }z}e^{-{|{\lambda }|^2\over 2}}\), the density matrix elements can be written as

$$\begin{aligned} (\,\psi _t\,,a_i^+a_j \,\psi _t\,)_\mathcal{F}= & {} \textsf{E}_x\bar{\textsf{E}}_y\bigl [ \;v_{i,x,t} \;{\bar{v}}_{j,y,t}\; e^{v_{x,t}{\bar{v}}_{y,t}} \;\bigr ] \;e^{-|{\lambda }|^2} \end{aligned}$$
(2.29)

with vector valued functions \(\,v=v_{x,t}\in \mathbb C^{|\Gamma |}\,\) and \(\,{\bar{v}}={\bar{v}}_{y,t}\in \mathbb C^{|\Gamma |}\,\) which are given by the SDE system

$$\begin{aligned} dv= & {} \;-\;i\,\bigl (\, dt\,\varepsilon \;+\;\sqrt{2u}\,dx_t\,\bigr )\, v \nonumber \\ d\bar{v}= & {} \;+\;i\,\bigl (\, dt\,\varepsilon \;+\;\sqrt{2u}\,dy_t\,\bigr )\, {\bar{v}}\end{aligned}$$
(2.30)

with initial conditions \(\,v_{x,0}={\lambda },\,{\bar{v}}_{y,0}=\bar{\lambda }\,\). Since \(\,dx_{j,t}\,dy_{j,t}=0\,\), we obtain

$$\begin{aligned} d(v^T{\bar{v}})= & {} dv^T{\bar{v}}\;+\;v^Td{\bar{v}}\;+\;dv^Td{\bar{v}}\nonumber \\= & {} \;-\;i\,v^T\bigl (\, dt\,\varepsilon \;+\;\sqrt{2u} \,dx_{t}\,\bigr ){\bar{v}}\; +\; i\, v^T\bigl (\, dt\,\varepsilon \;+\;\sqrt{2u} \,dy_{t}\,\bigr )\,{\bar{v}}\;\;+\;0 \nonumber \\= & {} \;-\;i\sqrt{2u}\, v^T(dx_t-dy_t) {\bar{v}}\nonumber \\= & {} \;-\;i\sqrt{2u}\, {\sum }_j v_j{\bar{v}}_j\,(dx_{j,t}-dy_{j,t}) \end{aligned}$$
(2.31)

That is, the quantity \(\,v^T{\bar{v}}={\sum }_{j} v_j{\bar{v}}_j\,\equiv \,v{\bar{v}}\,\), we omit the transpose sign in the following, is a martingale. Now we write

$$\begin{aligned} (v{\bar{v}})_{kdt}= & {} (v{\bar{v}})_0 \;+\; \sum _{\ell =1}^k \bigl [\,(v{\bar{v}})_{\ell dt} \,-\,(v{\bar{v}})_{(\ell -1) dt}\,\bigr ] \nonumber \\= & {} (v{\bar{v}})_0 \;+\; \sum _{\ell =1}^k d(v{\bar{v}})_{\ell dt} \nonumber \\= & {} (v{\bar{v}})_0 \;-\;i\sqrt{2u}\, \sum _{\ell =1}^k{\sum }_j (v_j{\bar{v}}_j)_{(\ell -1)dt} \,(dx_{j,\ell dt} - dy_{j,\ell dt}) \nonumber \\= & {} |{\lambda }|^2 \;-\;i\sqrt{2udt}\; \sum _{\ell =1}^k{\sum }_j (v_j{\bar{v}}_j)_{(\ell -1)dt} \,(\phi _{j,\ell } - \theta _{j,\ell }) \end{aligned}$$
(2.32)

such that

$$\begin{aligned} e^{(v{\bar{v}})_{kdt}}\,e^{-|{\lambda }|^2}\;\;=\;\;e^{(v{\bar{v}})_{kdt}-(v{\bar{v}})_0}= & {} \exp \biggl \{\;-\;i\sqrt{2udt}\; \sum _{\ell =1}^k\sum _j (v_j{\bar{v}}_j)_{(\ell -1)dt} \,(\phi _{j,\ell } - \theta _{j,\ell }) \;\biggr \} \end{aligned}$$
(2.33)

with integration variables

$$\begin{aligned} dx_{j,\ell dt}= & {} \sqrt{dt}\,\phi _{j,\ell } \\ dy_{j,\ell dt}= & {} \sqrt{dt}\,\theta _{j,\ell } \nonumber \end{aligned}$$
(2.34)

In terms of the \(\,\phi _{j,\ell },\theta _{j,\ell }\,\) variables, the Fresnel measure reads

$$\begin{aligned} \textsf{E}_x\bar{\textsf{E}}_y\bigl [\;\cdots \;\bigr ]= & {} \int _{{\mathbb R}^{k{|\Gamma |}}}\int _{{\mathbb R}^{k{|\Gamma |}}} \;\;\cdots \;\;\; \mathop \Pi _{\ell =1}^k \; e^{ {i\over 2}\sum _j (\phi _{j,\ell }^2 - \theta _{j,\ell }^2) } \; { d^{|\Gamma |}\phi _\ell \,d^{|\Gamma |}\theta _\ell \over (2\pi )^{|\Gamma |}} \end{aligned}$$
(2.35)

Consider the \(\ell \)’th term on the right hand side of (2.33),

$$\begin{aligned} \sum _j (v_j{\bar{v}}_j)_{(\ell -1)dt} \,(\phi _{j,\ell } - \theta _{j,\ell }) \end{aligned}$$
(2.36)

The \(\phi _\ell \) and \(\theta _\ell \) show up in an explicit linear form, since the quantities

$$\begin{aligned} (v_j{\bar{v}}_j)_{(\ell -1)dt}= & {} (v_j{\bar{v}}_j)_{(\ell -1)dt}\Bigl (\, \{\phi _m,\theta _m\}_{m=1}^{\ell -1}\, \Bigr ) \end{aligned}$$
(2.37)

depend only on \(\phi \)’s and \(\theta \)’s at earlier times \(t_1,\cdots ,t_{\ell -1}\). Thus, we can absorb them into the integration measure simply by completing the square. In the mathematics literature, the corresponding change of variables then is called a Girsanov transformation. Thus, this is a very elementary calculation, but, since this is a key step, let us be very explicit and proceed line by line. The result is summarized in Theorem 4 below.

We have at time \(\,t=t_k=kdt\)

$$\begin{aligned} { (\,\psi _t\,,a_i^+a_j \,\psi _t\,)_\mathcal{F}\;\;=\;\;\textsf{E}\bar{\textsf{E}}[ \,v_i {\bar{v}}_j\, e^{v{\bar{v}}} \,] \;e^{-|{\lambda }|^2}} \\= & {} \int _{{\mathbb R}^{2k{|\Gamma |}}} v_{i,kdt}\, {\bar{v}}_{j,kdt}\; \prod _{\ell =1}^k \exp \biggl \{\;-\;i\sqrt{2udt}\;\sum _j (v_j{\bar{v}}_j)_{(\ell -1)dt} \,(\phi _{j,\ell } - \theta _{j,\ell }) \;\biggr \} \;\times \nonumber \\{} & {} \prod _{\ell =1}^k \; \exp \biggl \{\, {i\over 2}\sum _j (\phi _{j,\ell }^2 - \theta _{j,\ell }^2) \,\biggr \} \; { d^{|\Gamma |}\phi _\ell \, d^{|\Gamma |}\theta _\ell \over (2\pi )^{|\Gamma |}} \nonumber \\ \nonumber \end{aligned}$$
(2.38)

Consider the \(\ell \)-th factor. The exponentials with the \(\phi _\ell \) variables combine to

$$\begin{aligned} { \exp \biggl \{\;-\;i\sqrt{2udt}\;\sum _j (v_j{\bar{v}}_j)_{(\ell -1)dt} \,\phi _{j,\ell } \;\biggr \} \;\times \; \exp \biggl \{\, {i\over 2}\sum _j \phi _{j,\ell }^2 \,\biggr \}} \\= & {} \exp \biggl \{\; {i\over 2}\sum _j \Bigl [\, \phi _{j,\ell }^2 \;-\; 2\sqrt{2udt}\, (v_j{\bar{v}}_j)_{(\ell -1)dt} \,\phi _{j,\ell } \,\Bigr ] \; \biggr \} \nonumber \\= & {} \exp \biggl \{\; {i\over 2}\sum _j \bigl [\, \phi _{j,\ell } \,-\, \sqrt{2udt}\, (v_j{\bar{v}}_j)_{(\ell -1)dt} \,\bigr ]^2 \; \biggr \} \;\times \; \exp \biggl \{\;-\, {i\over 2}\sum _j \, 2udt\, (v_j{\bar{v}}_j)_{(\ell -1)dt}^2 \; \biggr \} \nonumber \end{aligned}$$
(2.39)

The exponentials with the \(\theta _\ell \) variables combine to

$$\begin{aligned} { \exp \biggl \{\;+\;i\sqrt{2udt}\;\sum _j (v_j{\bar{v}}_j)_{(\ell -1)dt} \,\theta _{j,\ell } \;\biggr \} \;\times \; \exp \biggl \{\,-\, {i\over 2}\sum _j \theta _{j,\ell }^2 \,\biggr \}} \\= & {} \exp \biggl \{\;-\, {i\over 2}\sum _j \Bigl [\, \theta _{j,\ell }^2 \;-\; 2\sqrt{2udt}\, (v_j{\bar{v}}_j)_{(\ell -1)dt} \,\theta _{j,\ell } \,\Bigr ] \; \biggr \} \nonumber \\= & {} \exp \biggl \{\;-\, {i\over 2}\sum _j \bigl [\, \theta _{j,\ell } \,-\, \sqrt{2udt}\, (v_j{\bar{v}}_j)_{(\ell -1)dt} \,\bigr ]^2 \; \biggr \} \;\times \; \exp \biggl \{\;+\, {i\over 2}\sum _j \, 2udt\, (v_j{\bar{v}}_j)_{(\ell -1)dt}^2 \; \biggr \} \nonumber \\ \nonumber \end{aligned}$$
(2.40)

Observe that the last exponentials in (2.39) and (2.40)

$$\begin{aligned} \exp \bigl \{\;\!-\!\, {i\over 2}{\sum }_j \, 2udt\, (v_j{\bar{v}}_j)_{(\ell -1)dt}^2 \; \bigr \}\times \exp \bigl \{\;\!+\!\, {i\over 2}{\sum }_j \, 2udt\, (v_j{\bar{v}}_j)_{(\ell -1)dt}^2 \; \bigr \}= & {} 1 \end{aligned}$$
(2.41)

cancel each other.

Now we make the substitution of variables

$$\begin{aligned} \tilde{\phi }_{j,\ell }:= & {} \phi _{j,\ell } \,-\, \sqrt{2udt}\, (v_j{\bar{v}}_j)_{(\ell -1)dt} \\ \tilde{\theta }_{j,\ell }:= & {} \theta _{j,\ell } \,-\, \sqrt{2udt}\, (v_j{\bar{v}}_j)_{(\ell -1)dt} \nonumber \end{aligned}$$
(2.42)

or equivalently

$$\begin{aligned} d\tilde{x}_{j,t_\ell }:= & {} dx_{j,t_\ell } \,-\, \sqrt{2u}\,dt\, (v_j{\bar{v}}_j)_{(\ell -1)dt} \\ d\tilde{y}_{j,t_\ell }:= & {} dy_{j,t_\ell } \,-\, \sqrt{2u}\,dt\, (v_j{\bar{v}}_j)_{(\ell -1)dt} \nonumber \end{aligned}$$
(2.43)

Then we can write

$$\begin{aligned} (\,\psi _t\,,a_i^+a_j \,\psi _t\,)_\mathcal{F}= & {} \textsf{E}_x\bar{\textsf{E}}_y[ \,v_i {\bar{v}}_j\, e^{v{\bar{v}}} \,] \;e^{-|{\lambda }|^2} \nonumber \\= & {} \textsf{E}_{\tilde{x}}\bar{\textsf{E}}_{\tilde{y}}[ \,v_i {\bar{v}}_j\, ] \end{aligned}$$
(2.44)

where in terms of the transformed variables \(\tilde{x},\tilde{y}\,\) the \(v_j\) and \({\bar{v}}_j\) are given by the transformed SDE system

$$\begin{aligned} dv_j= & {} \;-\;i\, dt\,(\varepsilon v)_j \;-\;i\sqrt{2u}\,v_j\,dx_j \nonumber \\= & {} \;-\;i\, dt\,(\varepsilon v)_j \;-\;i\sqrt{2u}\,v_j\,\bigl [ d\tilde{x}_j \,+\, \sqrt{2u}\,dt\, v_j{\bar{v}}_j \bigr ] \nonumber \\= & {} \;-\;i\, dt\,(\varepsilon v)_j \;-\; i\,2u\,dt\, v_j{\bar{v}}_j \,v_j\;-\;i\sqrt{2u}\,v_j\, d\tilde{x}_j \end{aligned}$$
(2.45)

and

$$\begin{aligned} d\bar{v}_j= & {} \;+\;i\, dt\,(\varepsilon \bar{v})_j \;+\;i\sqrt{2u}\,\bar{v}_j\,dy_j \nonumber \\= & {} \;+\;i\, dt\,(\varepsilon \bar{v})_j \;+\;i\sqrt{2u}\,\bar{v}_j\,\bigl [ d\tilde{y}_j \,+\, \sqrt{2u}\,dt\, v_j{\bar{v}}_j \bigr ] \nonumber \\= & {} \;+\;i\, dt\,(\varepsilon \bar{v})_j \;+\;i\, 2u\,dt\, v_j{\bar{v}}_j \,{\bar{v}}_j \;+\;i\sqrt{2u}\,\bar{v}_j\, d\tilde{y}_j \end{aligned}$$
(2.46)

with initial conditions

$$\begin{aligned} v_{j,0}= & {} {\lambda }_j \nonumber \\ {\bar{v}}_{j,0}= & {} {\bar{\lambda }}_j \end{aligned}$$
(2.47)

Recall that

$$\begin{aligned} |{\lambda }|^2= & {} N \end{aligned}$$
(2.48)

has the meaning of total number of particles. Thus, if we devide (2.45) and (2.46) through \(|{\lambda }|=\sqrt{N}\) and put

$$\begin{aligned} w_j:= & {} v_j\;/\;|{\lambda }| \nonumber \\ \bar{w}_j:= & {} {\bar{v}}_j\;/\;|{\lambda }| \end{aligned}$$
(2.49)

we obtain

$$\begin{aligned} dw_j= & {} \;-\;i\, dt\,(\varepsilon w)_j \;-\; i\,2uN\,dt\, w_j\bar{w}_j \,w_j\;-\;i\sqrt{2u}\,w_j\, d\tilde{x}_j \nonumber \\ d\bar{w}_j= & {} \;+\;i\, dt\,(\varepsilon \bar{w})_j\;+\;i\, 2uN\,dt\, w_j\bar{w}_j \,\bar{w}_j \;+\;i\sqrt{2u}\,\bar{w}_j\, d\tilde{y}_j \end{aligned}$$
(2.50)

or, with

$$\begin{aligned} g:= & {} uN \end{aligned}$$
(2.51)
$$\begin{aligned} dw_j= & {} \;-\;i\, dt\,(\varepsilon w)_j \;-\; i\,2g\,dt\, w_j\bar{w}_j \,w_j\;-\;i\sqrt{2g/N}\,w_j\, d\tilde{x}_j \nonumber \end{aligned}$$
(2.52)
$$\begin{aligned} d\bar{w}_j= & {} \;+\;i\, dt\,(\varepsilon \bar{w})_j\;+\;i\, 2g\,dt\, w_j\bar{w}_j \,\bar{w}_j \;+\;i\sqrt{2g/N}\,\bar{w}_j\, d\tilde{y}_j \end{aligned}$$
(2.53)

Thus, in the limit \(\,N\rightarrow \infty \,\) with \(\,g=uN\,\) fixed, the diffusive part vanishes, the SDE reduces to a deterministic ODE system and the exact density matrix elements are given by

$$\begin{aligned} (\,\psi _t\,,a_i^+a_j \,\psi _t\,)_\mathcal{F}= & {} \textsf{E}_{\tilde{x}}\bar{\textsf{E}}_{\tilde{y}}[ \,v_i {\bar{v}}_j\, ] \;\;=\;\; N \;\times \; w_i(t) \,\bar{w}_j(t) \end{aligned}$$
(2.54)

with the \(w,\bar{w}\) given by the ODE system

$$\begin{aligned} \dot{w}_j= & {} \;-\;i\,(\varepsilon w)_j \;-\; i\,2g\, w_j\bar{w}_j \,w_j \nonumber \\ \dot{\bar{w}}_j= & {} \;+\;i\,(\varepsilon \bar{w})_j\;+\;i\, 2g\, w_j\bar{w}_j \,\bar{w}_j \end{aligned}$$
(2.55)

with initial conditions

$$\begin{aligned} w_j(0)= & {} {\lambda }_j\;/\;|{\lambda }| \nonumber \\ \bar{w}_j(0)= & {} {\bar{\lambda }}_j\;/\;|{\lambda }| \end{aligned}$$
(2.56)

Observe that now the \(\bar{w}_j\) are the true complex conjugates of the \(w_j\). Before we proceed to the analog calculation for number states in the next section, let’s summarize the results in the following

Theorem 4

Recall the SDE representation of Theorem 2 above,

$$\begin{aligned} (\,\psi _t\,,a_i^+a_j \,\psi _t\,)_\mathcal{F}= & {} \textsf{E}_x\bar{\textsf{E}}_y\bigl [ \;v_{i,x,t} \;{\bar{v}}_{j,y,t}\; e^{v_{x,t}{\bar{v}}_{y,t}} \;\bigr ] \;e^{-|{\lambda }|^2} \end{aligned}$$
(2.57)

with

$$\begin{aligned} dv_j= & {} \;-\;i\, dt\,(\varepsilon v)_j \;-\;i\,\sqrt{2u}\,v_j\,dx_j \nonumber \\ d\bar{v}_j= & {} \;+\;i\, dt\,(\varepsilon \bar{v})_j \;+\;i\,\sqrt{2u}\,\bar{v}_j\,dy_j \end{aligned}$$
(2.58)

Then, with the Girsanov transformation

$$\begin{aligned} d\tilde{x}_{j,t_\ell }:= & {} dx_{j,t_\ell } \,-\, \sqrt{2u}\,dt\, (v_j{\bar{v}}_j)_{(\ell -1)dt} \nonumber \\ d\tilde{y}_{j,t_\ell }:= & {} dy_{j,t_\ell } \,-\, \sqrt{2u}\,dt\, (v_j{\bar{v}}_j)_{(\ell -1)dt} \end{aligned}$$
(2.59)

the exponential \(\,e^{(v{\bar{v}})_t-|{\lambda }|^2}=e^{(v{\bar{v}})_t-(v{\bar{v}})_0}\,\) can be absorbed into the Fresnel integration measure, there is the identity

$$\begin{aligned} (\,\psi _t\,,a_i^+a_j \,\psi _t\,)_\mathcal{F}= & {} \textsf{E}_{\tilde{x}}\bar{\textsf{E}}_{\tilde{y}}\bigl [ \;v_{i,\tilde{x},t} \;{\bar{v}}_{j,\tilde{y},t}\; \bigr ] \end{aligned}$$
(2.60)

with

$$\begin{aligned} dv_j= & {} \;-\;i\, dt\,(\varepsilon v)_j \;-\; i\,2u\,dt\, v_j{\bar{v}}_j \,v_j\;-\;i\sqrt{2u}\,v_j\, d\tilde{x}_j \nonumber \\ d\bar{v}_j= & {} \;+\;i\, dt\,(\varepsilon \bar{v})_j\;+\;i\, 2u\,dt\, v_j{\bar{v}}_j \,{\bar{v}}_j \;+\;i\sqrt{2u}\,\bar{v}_j\, d\tilde{y}_j \end{aligned}$$
(2.61)

In the limit \(\,N\rightarrow \infty \,\) with \(\,g=uN\,\) fixed, we have the exact representation

$$\begin{aligned} (\,\psi _t\,,a_i^+a_j \,\psi _t\,)_\mathcal{F}= & {} N \;\times \; w_i(t) \,\bar{w}_j(t) \end{aligned}$$
(2.62)

with \(\bar{w}_j\) now being the true complex conjugate of \(w_j\) and the \(w_j\) are given by the ODE system

$$\begin{aligned} \dot{w}_j= & {} \;-\;i\,(\varepsilon w)_j \;-\; i\,2g\, w_j\bar{w}_j \,w_j \end{aligned}$$
(2.63)

with initial conditions \(\,w_j(0)= {\lambda }_j\,/\,|{\lambda }|\,\). Equation (2.62) is the time dependent discrete Gross Pitaevskii equation, here with a general hopping matrix \(\varepsilon =(\varepsilon _{ij})_{i,j\in \Gamma }\) which may also include some on-diagonal trapping potentials \(\epsilon _j=\varepsilon _{jj}\) from the Hamiltonian (1.1).

2.3 Number States

Let’s consider the dynamics of number states which are used to describe the dynamics of Bose-Einstein condensates. They are given by the following initial state

$$\begin{aligned} \psi _0(z)\;=\;\psi _0(\,\{z_j\}\,)= & {} {1 \over \sqrt{N!}\,|\lambda |^N} \;\bigl (\, {\sum }_j {\lambda }_j z_j \,\bigr )^N\;\;=\;\; {({\lambda }z)^N \over \sqrt{{N!}\,|\lambda |^N}} \end{aligned}$$
(2.64)

with

$$\begin{aligned} |{\lambda }|^2= & {} {\sum }_j |{\lambda }_j|^2\;\;=\;\; N \end{aligned}$$
(2.65)

There are the following formulae which are standard expectations over the bosonic Fock space, here written again in the Bargmann-Segal representation:

$$\begin{aligned} {1\over N! }\;\int _{\mathbb C^{|\Gamma |}}\, ({\lambda }z)^N\, ({\bar{\lambda }}{\bar{z}})^N \; d\mu (z)= & {} (\,|{\lambda }|^2\,)^N \end{aligned}$$
(2.66)
$$\begin{aligned} {1\over N!}\;\int _{\mathbb C^{|\Gamma |}}\, z_i\,\bar{z}_j\;({\lambda }z)^N\, ({\bar{\lambda }}{\bar{z}})^N \; d\mu (z)= & {} \delta _{i,j}\;(\,|{\lambda }|^2\,)^N \;+\;N\,{\bar{\lambda }}_i{\lambda }_j\;(\,|{\lambda }|^2\,)^{N-1} \end{aligned}$$
(2.67)

Thus, with the condition (2.64), we have the following time zero expectations:

$$\begin{aligned} (\psi _0,\psi _0)_\mathcal{F}= & {} \int _{\mathbb C^{|\Gamma |}} |\psi _0(z)|^2 \,d\mu (z) \;\;=\;\; { |{\lambda }|^{2N} \over N^N} \;\;=\;\;1 \end{aligned}$$
(2.68)
$$\begin{aligned} (\,\psi _0\,,a_i^+a_j \,\psi _0\,)_\mathcal{F}= & {} \int _{\mathbb C^{|\Gamma |}} z_j {\bar{z}}_i\; |\psi _0(z)|^2 \,d\mu (z) \;\;-\;\; \delta _{i,j} \;\;=\;\; N\;{{\lambda }_i{\bar{\lambda }}_j \over |{\lambda }|^2}\;\;=\;\;{\lambda }_i{\bar{\lambda }}_j \ \ \ \ \ \ \end{aligned}$$
(2.69)

Now, let’s consider the time t formulae. From (1.58) and (1.62) we have for an arbitrary initial state \(\psi _0\)

$$\begin{aligned} { (\,\psi _t\,,a_i^+a_j \,\psi _t\,)_\mathcal{F}\;\;=\;\; \int _{\mathbb C^{|\Gamma |}} z_j {\bar{z}}_i\; |\psi _t(z)|^2 \,d\mu (z) \;\;-\;\; \delta _{i,j} }\\= & {} \int \int \;\Bigl \{\; \int _{\mathbb C^{|\Gamma |}} z_j {\bar{z}}_i\; \psi _0(U_{x,t}z)\;\overline{\psi _0(U_{y,t}z)}\;d\mu (z)\;\Bigr \} \; dF(\{x_t\})\; d\bar{F}(\{y_t\}) \;\;-\;\; \delta _{i,j} \nonumber \end{aligned}$$
(2.70)

The wavy bracket we can evaluate with the formulae from above. We obtain

$$\begin{aligned} { \int _{\mathbb C^{|\Gamma |}} z_j {\bar{z}}_i\; \psi _0(U_{x,t}z)\;\overline{\psi _0(U_{y,t}z)}\;d\mu (z) } \\= & {} {1\over N!N^N}\;\int _{\mathbb C^{|\Gamma |}}\, z_j\,\bar{z}_i\;\bigl (\,U_{x,t}^T{\lambda }\cdot z\,\bigr )^N\, \bigl (\,\bar{U}_{y,t}^T{\lambda }\cdot {\bar{z}}\,\bigr )^N \; d\mu (z) \nonumber \\= & {} {1\over N^N}\;\Bigl \{\; \delta _{i,j}\;(\,{\lambda }\cdot U_{x,t}U_{y,t}^+{\bar{\lambda }}\,)^N \;+\;N\,[U_{x,t}^T{\lambda }]_i\; [\bar{U}_{y,t}^T{\bar{\lambda }}]_j\;(\,{\lambda }\cdot U_{x,t}U_{y,t}^+{\bar{\lambda }}\,)^{N-1} \;\Bigr \} \nonumber \end{aligned}$$
(2.71)

Thus,

$$\begin{aligned} { (\,\psi _0\,,a_i^+a_j \,\psi _0\,)_\mathcal{F}\;\;=\;\; } \\{} & {} {1\over N^N}\;\textsf{E}_x\bar{\textsf{E}}_y\Bigl [\; \delta _{i,j}\;(\,{\lambda }\cdot U_{x,t}U_{y,t}^+{\bar{\lambda }}\,)^N \;+\;N\,[U_{x,t}^T{\lambda }]_i\; [\bar{U}_{y,t}^T{\bar{\lambda }}]_j\;(\,{\lambda }\cdot U_{x,t}U_{y,t}^+{\bar{\lambda }}\,)^{N-1} \;\Bigr ] \;-\;\delta _{i,j} \nonumber \\= & {} \textsf{E}_x\bar{\textsf{E}}_y\Bigl [\; [U_{x,t}^T{\lambda }]_i\; [\bar{U}_{y,t}^T{\bar{\lambda }}]_j \;\bigl (\, U_{x,t}^T{\lambda }\,\cdot \, U_{y,t}^+{\bar{\lambda }}\,\bigr )^{N-1} \;\Bigr ] \;\bigr /\;({\lambda }\cdot {\bar{\lambda }})^{N-1} \nonumber \end{aligned}$$
(2.72)

since

$$\begin{aligned} {1\over N^N}\;\textsf{E}_x\bar{\textsf{E}}_y\Bigl [\; \delta _{i,j}\;(\,{\lambda }\cdot U_{x,t}U_{y,t}^+{\bar{\lambda }}\,)^N \;\Bigr ]= & {} \delta _{i,j} \end{aligned}$$
(2.73)

because of Theorem 3 of Section 2.1. Now we can proceed as in the preceeding Section 2.2. We introduce the variables

$$\begin{aligned} v_j\;=\;v_{x,t,j}:= & {} [\,U_{x,t}^T{\lambda }\,]_{j} \nonumber \\ {\bar{v}}_j\;=\;{\bar{v}}_{y,t,j}:= & {} [\,\bar{U}_{y,t}^T{\bar{\lambda }}\,]_{j} \end{aligned}$$
(2.74)

and from the matrix equations

$$\begin{aligned} dU_{x,t}^T= & {} -i\,\bigl (\, dt\,\varepsilon \;+\;\sqrt{2u} \,dx_{t}\,\bigr )\,U_{x,t}^T \nonumber \\ d{\bar{U}}_{y,t}^T= & {} +i\,\bigl (\, dt\,\varepsilon \;+\;\sqrt{2u} \,dy_{t}\,\bigr )\,{\bar{U}}_{y,t}^T \end{aligned}$$
(2.75)

we obtain

$$\begin{aligned} dv_j= & {} -i\, dt\,(\varepsilon v)_j\;+\;\sqrt{2u} \,dx_{j,t}\,v_j \nonumber \\ d{\bar{v}}_j= & {} +i\, dt\,(\varepsilon {\bar{v}})_j\;+\;\sqrt{2u} \,dy_{j,t}\,{\bar{v}}_j \end{aligned}$$
(2.76)

Furthermore,

$$\begin{aligned} \bigl (\, U_{x,t}^T{\lambda }\,\cdot \, U_{y,t}^+{\bar{\lambda }}\,\bigr )^{N-1}= & {} \bigl (\, {{\sum }}_{j}{v_j}{\bar{v}}_j\,\bigr )^{N-1}\;\;=:\;\; (v{\bar{v}})^{N-1} \end{aligned}$$
(2.77)

Thus, for the density matrix elements we obtain the following representation:

$$\begin{aligned} (\,\psi _t\,,a_i^+a_j \,\psi _t\,)_\mathcal{F}= & {} \textsf{E}\bar{\textsf{E}}\bigl [ \, v_i\,{\bar{v}}_j\; (v{\bar{v}})^{N-1} \,\bigr ] \;\bigr /\;({\lambda }{\bar{\lambda }})^{N-1} \end{aligned}$$
(2.78)

As in the last section, we can make a Girsanov transformation and absorb the quantity \(\,(v{\bar{v}})^{N-1}\,\) into the Fresnel measure. Since the \(v,{\bar{v}}\) obey exactly the same SDEs as in the last section with the same initial conditions

$$\begin{aligned} v_{0}= & {} {\lambda }\nonumber \\ {\bar{v}}_{0}= & {} {\bar{\lambda }}\end{aligned}$$
(2.79)

Equation (2.31) remains unchanged:

$$\begin{aligned} d(v{\bar{v}})= & {} \;-\;i\sqrt{2u}\, {{\sum }}_{j} v_j{\bar{v}}_j\,(dx_{j,t}-dy_{j,t}) \end{aligned}$$
(2.80)

Now we abbreviate

$$\begin{aligned} P(v{\bar{v}}):= & {} (v{\bar{v}})^{N-1} \end{aligned}$$
(2.81)
$$\begin{aligned} p(v{\bar{v}}):= & {} P'(v{\bar{v}})/P(v{\bar{v}})\;=\;[\log P]'(v{\bar{v}})\;=\; (N-1)/(v{\bar{v}}) \end{aligned}$$
(2.82)

Then, since \((v{\bar{v}})_0={\lambda }{\bar{\lambda }}\) and, more importantly, \([d(v{\bar{v}})]^2=0\), we can write

$$\begin{aligned} {(v_t{\bar{v}}_t)^{N-1}\over ({\lambda }{\bar{\lambda }})^{N-1}}= & {} e^{\log P(v_t{\bar{v}}_t)-\log P(v_0{\bar{v}}_0)}\;\;=\;\;e^{ \int _0^t p(v_s{\bar{v}}_s)\,d(v{\bar{v}})_s } \end{aligned}$$
(2.83)

With the discrete time variables of Section 2.2, the exponent reads as follows:

$$\begin{aligned} \int _0^{t_k} p(v_s{\bar{v}}_s)\,d(v{\bar{v}})_s= & {} {{\sum }}_{\ell =1}^k p[(v{\bar{v}})_{(\ell -1)dt}] \, d(v{\bar{v}})_{\ell dt} \nonumber \\= & {} \;-\;i\sqrt{2u}\, {\sum }_{\ell =1}^k p[(v{\bar{v}})_{(\ell -1)dt}]\, {\sum }_j (v_j{\bar{v}}_j)_{(\ell -1)dt} \,(dx_{j,\ell dt} - dy_{j,\ell dt}) \nonumber \\= & {} \;-\;i\sqrt{2udt}(N-1)\; {\sum }_{\ell =1}^k{\sum }_j \bigl ({v_j{\bar{v}}_j\over v{\bar{v}}}\bigr )_{(\ell -1)dt} \,(\phi _{j,\ell } - \theta _{j,\ell }) \end{aligned}$$
(2.84)

Thus, now we have to make the following substitution of variables:

$$\begin{aligned} \tilde{\phi }_{j,\ell }:= & {} \phi _{j,\ell } \,-\, \sqrt{2udt}(N-1)\, (v_j{\bar{v}}_j/v{\bar{v}})_{(\ell -1)dt} \nonumber \\ \tilde{\theta }_{j,\ell }:= & {} \theta _{j,\ell } \,-\, \sqrt{2udt}(N-1)\, (v_j{\bar{v}}_j/v{\bar{v}})_{(\ell -1)dt} \end{aligned}$$
(2.85)

or equivalently

$$\begin{aligned} d\tilde{x}_{j,t_\ell }:= & {} dx_{j,t_\ell } \,-\, \sqrt{2u}(N-1)\,dt\, (v_j{\bar{v}}_j/v{\bar{v}})_{(\ell -1)dt} \nonumber \\ d\tilde{y}_{j,t_\ell }:= & {} dy_{j,t_\ell } \,-\, \sqrt{2u}(N-1)\,dt\, (v_j{\bar{v}}_j/v{\bar{v}})_{(\ell -1)dt} \end{aligned}$$
(2.86)

Then we can write

$$\begin{aligned} (\,\psi _t\,,a_i^+a_j \,\psi _t\,)_\mathcal{F}= & {} \textsf{E}_x\bar{\textsf{E}}_y[ \,v_i {\bar{v}}_j\, (v{\bar{v}})^{N-1} \,]\,/\,({\lambda }{\bar{\lambda }})^{N-1} \;\;=\;\; \textsf{E}_{\tilde{x}}\bar{\textsf{E}}_{\tilde{y}}[ \,v_i {\bar{v}}_j\, ] \end{aligned}$$
(2.87)

where in terms of the transformed variables \(\tilde{x},\tilde{y}\,\) the \(v_j\) and \({\bar{v}}_j\) are given by the transformed SDE system

$$\begin{aligned} dv_j= & {} \;-\;i\, dt\,(\varepsilon v)_j \;-\;i\sqrt{2u}\,v_j\,dx_j \nonumber \\= & {} \;-\;i\, dt\,(\varepsilon v)_j \;-\; i\,2u(N-1)\,dt\, {v_j{\bar{v}}_j\over v{\bar{v}}} \,v_j\;-\;i\sqrt{2u}\,v_j\, d\tilde{x}_j \end{aligned}$$
(2.88)

and

$$\begin{aligned} d\bar{v}_j= & {} \;+\;i\, dt\,(\varepsilon \bar{v})_j \;+\;i\sqrt{2u}\,\bar{v}_j\,dy_j \nonumber \\= & {} \;+\;i\, dt\,(\varepsilon \bar{v})_j \;+\;i\, 2u(N-1)\,dt\, {v_j{\bar{v}}_j\over v{\bar{v}}} \,{\bar{v}}_j \;+\;i\sqrt{2u}\,\bar{v}_j\, d\tilde{y}_j \end{aligned}$$
(2.89)

with initial conditions \((v,{\bar{v}})_0=({\lambda },{\bar{\lambda }})\,\). Dividing the system through \(|{\lambda }|=\sqrt{N}\) and introducing again the normalized quantities

$$\begin{aligned} w:= & {} v\,/\,|{\lambda }| \nonumber \\ \bar{w}:= & {} {\bar{v}}\,/\,|{\lambda }| \end{aligned}$$
(2.90)

we obtain the following ODE system in the limit \(N\rightarrow \infty \,\) with \(g=UN\) fixed:

$$\begin{aligned} \dot{w}_j= & {} \;-\;i\,(\varepsilon w)_j \;-\; i\,2g\, {w_j\bar{w}_j\over w\bar{w}} \,w_j \nonumber \\ {\dot{\bar{w}}}_j= & {} \;+\;i\,(\varepsilon \bar{w})_j \;+\;i\,2g\, {w_j\bar{w}_j\over w\bar{w}} \,\bar{w}_j \end{aligned}$$
(2.91)

In particular, for a symmetric hopping matrix \(\,\varepsilon =\varepsilon ^T\)

$$\begin{aligned} {d\over dt}(w{\bar{w}})= & {} {\sum }_j\bigl \{\, \dot{w}_j\,{\bar{w}}_j\,+\,w_j\,{\dot{\bar{w}}}_j\,\bigr \} \nonumber \\= & {} \;-\;i\,{\sum }_j\bigl \{\,(\varepsilon w)_j\,{\bar{w}}_j\,-\,w_j\,(\varepsilon \bar{w})_j\,\bigr \} \;-\;i\,2g\,{\sum }_j\bigl \{\, {w_j{\bar{w}}_j\over w{\bar{w}}} \,w_j\,{\bar{w}}_j\,- \,w_j\, {w_j{\bar{w}}_j\over w{\bar{w}}} \,{\bar{w}}_j \,\bigr \} \nonumber \\= & {} \;\;0 \end{aligned}$$
(2.92)

Thus we have \((w{\bar{w}})_t=1\) for all t and the ODE system (2.90) reduces again to the time dependent discrete Gross-Pitaevskii equation. We summarize in the following

Theorem 5

Consider the following normalized initial number state,

$$\begin{aligned} \psi _0(z)\;=\;\psi _0(\,\{z_j\}\,)= & {} {1 \over \sqrt{N!}\,|\lambda |^N} \;\bigl (\, {\sum }_j {\lambda }_j z_j \,\bigr )^N\;\;=\;\; {({\lambda }z)^N \over \sqrt{N!}\,|\lambda |^N} \end{aligned}$$
(2.93)

with \(|{\lambda }|^2=N\). Then the time t density matrix elements can be written as

$$\begin{aligned} (\,\psi _t\,,a_i^+a_j \,\psi _t\,)_\mathcal{F}= & {} \textsf{E}\bar{\textsf{E}}[ \,v_i {\bar{v}}_j\, ] \end{aligned}$$
(2.94)

where the \(v_j,{\bar{v}}_j\) are given by the SDE system

$$\begin{aligned} dv_j= & {} \;-\;i\,(\varepsilon v)_j\, dt \;-\; i\,2u(N-1)\, {v_j{\bar{v}}_j\over v{\bar{v}}} \,v_j\,dt\;-\;i\sqrt{2u}\,v_j\, dx_j \nonumber \\ d\bar{v}_j= & {} \;+\;i\,(\varepsilon \bar{v})_j\, dt \;+\;i\, 2u(N-1)\, {v_j{\bar{v}}_j\over v{\bar{v}}} \,{\bar{v}}_j\,dt \;+\;i\sqrt{2u}\,\bar{v}_j\, dy_j \end{aligned}$$
(2.95)

with initial conditions \((v,{\bar{v}})_0=({\lambda },{\bar{\lambda }})\,\). In the lage N limit with \(g=uN\) fixed, this reduces again, as in Theorem 4 where the initial state was a coherent state, to the time dependent discrete Gross-Pitaevskii equation

$$\begin{aligned} \dot{w}_j= & {} \;-\;i\,(\varepsilon w)_j \;-\; i\,2g\, {w_j{\bar{w}}_j} \,w_j \nonumber \\ {\dot{\bar{w}}}_j= & {} \;+\;i\,(\varepsilon \bar{w})_j \;+\;i\,2g\, {w_j{\bar{w}}_j} \,{\bar{w}}_j \end{aligned}$$
(3.1)

with normalized quantities \((w,{\bar{w}}):=(v,{\bar{v}})/|{\lambda }|\), initial conditions \((w,{\bar{w}})_0=({\lambda },{\bar{\lambda }})/|{\lambda }|\,\) and density matrix elements given by \((\psi _{t}\,a_{i}^+{a} a_{j} \psi _{t})_\mathcal{F}\,=\, N\,w_{i,t} {\bar{w}}_{j,t}\).

The results of Theorems 4 and 5 are in line with rigorous results in the continuous case in the large N limit. In [2], Benedikter, Porta and Schlein give an overview on rigorous derivations of effective evolution equations and results concerning the continuous time dependent Gross-Pitaevskii equation are summarized in chapter 5. The article [3] focusses solely on the GP equation. Pickl [4, 5] and more recently Jeblick, Leopold and Pickl [6] also gave rigorous derivations of the continuous time dependent GP equation. The issue has a longer history with more people involved, more background can be found in [2]. The fact that the coherent states and the number states of Theorem 4 and 5 show similar dynamics in the large N limit has also been observed by Schachenmayer, Daley and Zoller in [7].

3 PDE Representations

3.1 Untransformed Case, before Girsanov Transformation

In the untransformed case, the SDE representation for the density matrix elements is given by Theorem 2 of Section 1.2. We have

$$\begin{aligned} (\,\psi _t\,,a_i^+a_j \,\psi _t\,)_\mathcal{F}= & {} \textsf{E}\bar{\textsf{E}}[ \,v_i {\bar{v}}_j\, e^{v{\bar{v}}} \,] \;e^{-|{\lambda }|^2} \end{aligned}$$
(3.2)

with

$$\begin{aligned} dv_j= & {} \;-\;i\, dt\,(\varepsilon v)_j \;-\;i\,\sqrt{2u}\,v_j\,dx_j \\ d\bar{v}_j= & {} \;+\;i\, dt\,(\varepsilon \bar{v})_j \;+\;i\,\sqrt{2u}\,\bar{v}_j\,dy_j \nonumber \end{aligned}$$
(3.3)

According to the Fresnel version of Kolmogorov’s backward equation, a one dimensional version is given in formula (A.81) with Fresnel expectation (A.74) in Appendix A.4, the quantity (3.1) has a PDE representation. In order to write it down, we need the operator A which is associated with the SDE system (3.2). To this end, we consider some arbitrary complex-valued function f of \(\,2{|\Gamma |}\) arguments,

$$\begin{aligned} f\;=\;f(\,\{v_j\},\{\bar{v}_j\} \,)&:&\mathbb C^{2{|\Gamma |}} \; \rightarrow \;\mathbb C \end{aligned}$$
(3.4)

Because of

$$\begin{aligned} (dv_j)^2= & {} \bigl \{\,-\;i\, dt(\varepsilon v)_j \;-\;i\sqrt{2u}\,v_j\,dx_j \,\bigr \}^2 \nonumber \\= & {} \bigl \{\,-\;i\sqrt{2u}\,v_j\,dx_j \,\bigr \}^2 \nonumber \\= & {} \;-\;2u\,v_j^2\,(dx_j)^2 \nonumber \\= & {} \;-\;i\,2u\,dt\, v_j^2 \end{aligned}$$
(3.5)

and

$$\begin{aligned} (d{\bar{v}}_j)^2= & {} \bigl \{\,+\;i\, dt(\varepsilon {\bar{v}})_j \;+\;i\sqrt{2u}\,{\bar{v}}_j\,dy_j \,\bigr \}^2 \nonumber \\= & {} \bigl \{\,+\;i\sqrt{2u}\,{\bar{v}}_j\,dy_j \,\bigr \}^2 \nonumber \\= & {} \;-\;2u\,{\bar{v}}_j^2\,(dy_j)^2 \nonumber \\= & {} \;+\;i\,2u\,dt\,{\bar{v}}_j^2 \end{aligned}$$
(3.6)

and, for \(i\ne j\),

$$\begin{aligned} dv_i \,d{\bar{v}}_i\;=\; dv_i\, d{\bar{v}}_j\;=\;dv_i \,dv_j\;=\;d{\bar{v}}_i \,d{\bar{v}}_j\;=\; 0 \end{aligned}$$
(3.7)

we obtain with the Fresnel version of the Ito lemma

$$\begin{aligned} df= & {} {\sum }_j { \Bigl \{\, {\partial f\over \partial v_j}\, dv_j \,+\, {\partial f\over \partial \bar{v}_j}\, d\bar{v}_j \,\Bigr \} } \;+\;{ {1\over 2}}\,{\sum }_j \Bigl \{\, {\partial ^2 f\over \partial v_j^2}\, (dv_j)^2 \,+\, {\partial ^2 f\over \partial \bar{v}_j^2}\, (d\bar{v}_j)^2 \,\Bigr \} \nonumber \\= & {} \;-\;i \,dt\;{\sum }_j { \Bigl \{\,(\varepsilon v)_j {\partial f\over \partial v_j}\,-\,(\varepsilon \bar{v})_j {\partial f\over \partial \bar{v}_j} \,\Bigr \} } \nonumber \\{} & {} \;-\;iu\,dt\;{\sum }_j \Bigl \{\,v_j^2\, {\partial ^2 f\over \partial v_j^2} \,-\,\bar{v}_j^2 \, {\partial ^2 f\over \partial \bar{v}_j^2}\,\Bigr \} \;\;\;+\;\;\textrm{diffusive} \nonumber \\=: & {} Af \;\;\;+\;\;\textrm{diffusive} \end{aligned}$$
(3.8)

Thus, the expectation

$$\begin{aligned} F_t:= & {} \textsf{E}\bar{\textsf{E}}\Bigl [ \,f(\,\{v_{j,x,t}\},\{\bar{v}_{j,y,t}\} \,)\;\Bigr ] \end{aligned}$$
(3.9)

considered as a function of its initial values \(\,(v_0,{\bar{v}}_0)=(\,\{v_{j,0}\},\{\bar{v}_{j,0}\} \,)=({\lambda },\bar{\lambda })\,\),

$$\begin{aligned} F_t= & {} F_t\bigl (\,\{v_{j,0}\},\{\bar{v}_{j,0}\} \,\bigr ) \end{aligned}$$
(3.10)

can be obtained as the solution of the parabolic second order PDE (the zero subscripts on the v’s are then usually omitted in the notation, \(v_{j,0}\rightarrow v_j\,\) in the following PDE)

$$\begin{aligned} {\partial F\over \partial t}\;=\;AF= & {} -i\,{\sum }_j { \Bigl \{\,(\varepsilon v)_j\, {\partial F\over \partial v_j}\,-\, (\varepsilon \bar{v})_j \,{\partial F\over \partial \bar{v}_j}\,\Bigr \} } \;-\;iu\,{\sum }_j \Bigl \{\; v_j^2 \,{\partial ^2 F\over \partial v_j^2}\;-\; \bar{v}_j^2 \,{\partial ^2 F\over \partial \bar{v}_j^2}\;\Bigr \} \end{aligned}$$
(3.11)

with initial condition

$$\begin{aligned} F_0= & {} f(\,\{v_j\},\{\bar{v}_j\}\,) \end{aligned}$$
(3.12)

If we introduce the differential operators

$$\begin{aligned} \mathcal{L}= & {} \mathcal{L}_0\;+\; \mathcal{L}_\textrm{int}\end{aligned}$$
(3.13)

with

$$\begin{aligned} \mathcal{L}_0:= & {} {\sum }_j \Bigl \{\; { (\varepsilon v)_j\, {\partial \over \partial v_j} \,-\, (\varepsilon \bar{v})_j\, {\partial \over \partial \bar{v}_j} } \;\Bigr \} \end{aligned}$$
(3.14)
$$\begin{aligned} \mathcal{L}_\textrm{int}:= & {} u\;{\sum }_j \Bigl \{\; v_j^2\, {\partial ^2 \over \partial v_j^2} \,-\,\bar{v}_j^2\, {\partial ^2 \over \partial \bar{v}_j^2} \;\Bigr \} \end{aligned}$$
(3.15)

the solution may be written as

$$\begin{aligned} F_t= & {} e^{-it\mathcal{L}}\,F_0 \end{aligned}$$
(3.16)

which one could try to evaluate through small or approximate large Trotter steps according to

$$\begin{aligned} F_{t}\;\;=\;\;e^{-it(\mathcal{L}_0+\mathcal{L}_\textrm{int})}\,F_0\approx & {} \bigl ( e^{-i{t\over k}t\mathcal{L}_0} \, e^{-i{t\over k}\mathcal{L}_\textrm{int}} \bigr )^k \,F_0 \end{aligned}$$
(3.17)

For the density matrix elements, we have to calculate the expectation (3.1) and hence the initial condition \(F_0=f\) is given by

$$\begin{aligned} F_0(\,\{v_j\},\{\bar{v}_j\}\,)= & {} v_i\,{\bar{v}}_j\; e^{\,v{\bar{v}}\,-\,|{\lambda }|^2} \end{aligned}$$
(3.18)

The time evolution is then obtained through

$$\begin{aligned} (\,\psi _t\,,a_i^+a_j \,\psi _t\,)_\mathcal{F}= & {} e^{-it(\mathcal{L}_0+\mathcal{L}_\textrm{int})}\,\bigl \{\, v_i\,{\bar{v}}_j\; e^{\,v{\bar{v}}\,-\,|{\lambda }|^2} \,\bigr \}\, \bigr |_{v={\lambda },{\bar{v}}={\bar{\lambda }}} \end{aligned}$$
(3.19)

In particular, the solution of the PDE (3.10) is not needed in the whole \((v,{\bar{v}})\) - space, but only at one specific point

$$\begin{aligned} (\,\{v_j\},\{\bar{v}_j\}\,)= & {} (\,\{{\lambda }_j\},\{\bar{\lambda }_j\}\,)\;\;. \end{aligned}$$
(3.20)

3.2 Transformed Case, after Girsanov Transformation

After Girsanov transformation, we have the SDE representation which is given by Theorem 4 of Section 2.2,

$$\begin{aligned} (\,\psi _t\,,a_i^+a_j \,\psi _t\,)_\mathcal{F}= & {} \textsf{E}\bar{\textsf{E}}[ \,v_i {\bar{v}}_j\, ] \end{aligned}$$
(3.21)

with

$$\begin{aligned} dv_j= & {} \;-\;i\, dt\,(\varepsilon v)_j \;-\; i\,2u\,dt\, v_j{\bar{v}}_j \,v_j \;-\;i\,\sqrt{2u}\,v_j\,dx_j \\ d\bar{v}_j= & {} \;+\;i\, dt\,(\varepsilon \bar{v})_j \;+\;i\, 2u\,dt\, v_j{\bar{v}}_j \,{\bar{v}}_j \;+\;i\,\sqrt{2u}\,\bar{v}_j\,dy_j \nonumber \end{aligned}$$
(3.22)

We still have

$$\begin{aligned} (dv_j)^2= & {} \;-\;i\,2u\,dt\, v_j^2 \\ (d{\bar{v}}_j)^2= & {} \;+\;i\,2u\,dt\,{\bar{v}}_j^2 \nonumber \end{aligned}$$
(3.23)

and, for \(i\ne j\),

$$\begin{aligned} dv_i \,d{\bar{v}}_i\;=\; dv_i\, d{\bar{v}}_j\;=\;dv_i \,dv_j\;=\;d{\bar{v}}_i \,d{\bar{v}}_j\;=\; 0 \end{aligned}$$
(3.24)

Thus, again with the Ito lemma, for some arbitrary \(\,f=f(\,\{v_j\},\{\bar{v}_j\} \,)\,\),

$$\begin{aligned} df= & {} {\sum }_j { \Bigl \{\, {\partial f\over \partial v_j}\, dv_j \,+\, {\partial f\over \partial \bar{v}_j}\, d\bar{v}_j \,\Bigr \} } \;+\;{ {1\over 2}}\,{\sum }_j \Bigl \{\, {\partial ^2 f\over \partial v_j^2}\, (dv_j)^2 \,+\, {\partial ^2 f\over \partial \bar{v}_j^2}\, (d\bar{v}_j)^2 \,\Bigr \} \nonumber \\= & {} \;-\;i \,dt\;{\sum }_j { \Bigl \{\,(\varepsilon v)_j {\partial f\over \partial v_j}\,-\,(\varepsilon \bar{v})_j {\partial f\over \partial \bar{v}_j} \,\Bigr \} } \;-\;i\,2u \,dt\;{\sum }_j { \Bigl \{\, v_{j}{\bar{v}}_{j}\,v_j \, {\partial f\over \partial v_j}\,-\,v_{j}{\bar{v}}_{j}\,{\bar{v}}_j \, {\partial f\over \partial \bar{v}_j} \,\Bigr \} } \nonumber \\{} & {} \;-\;iu\,dt\;{\sum }_j \Bigl \{\,v_j^2\, {\partial ^2 f\over \partial v_j^2} \,-\,\bar{v}_j^2 \, {\partial ^2 f\over \partial \bar{v}_j^2}\,\Bigr \} \;\;\;+\;\;\textrm{diffusive} \end{aligned}$$
(3.25)

such that the expectation

$$\begin{aligned} F_t:= & {} \textsf{E}\bar{\textsf{E}}\Bigl [ \,f(\,\{v_{j,x,t}\},\{\bar{v}_{j,y,t}\} \,)\;\Bigr ] \end{aligned}$$
(3.26)

considered again as a function of its initial values \(\,(v_0,{\bar{v}}_0)=(\,\{v_{j,0}\},\{\bar{v}_{j,0}\} \,)=({\lambda },\bar{\lambda })\,\),

$$\begin{aligned} F_t= & {} F_t\bigl (\,\{v_{j,0}\},\{\bar{v}_{j,0}\} \,\bigr ) \end{aligned}$$
(3.27)

and omitting the subscripts 0 on the \(v,{\bar{v}}\) in the following, can be obtained as the solution of the second order PDE

$$\begin{aligned} {\partial F\over \partial t}= & {} -i\,{\sum }_j { \Bigl \{\,(\varepsilon v)_j\, {\partial F\over \partial v_j}\,-\, (\varepsilon \bar{v})_j \,{\partial F\over \partial \bar{v}_j}\,\Bigr \} } \;-\;i\,2u \,{\sum }_j v_{j}{\bar{v}}_{j}\, { \Bigl \{\, v_j \, {\partial F\over \partial v_j}\,-\,{\bar{v}}_j \, {\partial F\over \partial \bar{v}_j} \,\Bigr \}} \nonumber \\{} & {} \;-\;i\,u\,{\sum }_j \Bigl \{\; v_j^2 \,{\partial ^2 F\over \partial v_j^2}\;-\; \bar{v}_j^2 \,{\partial ^2 F\over \partial \bar{v}_j^2}\;\Bigr \} \end{aligned}$$
(3.28)

with initial condition

$$\begin{aligned} F_0= & {} f(\,\{v_j\},\{\bar{v}_j\}\,) \end{aligned}$$
(3.29)

We write \(\,\mathcal{L}\,=\, \mathcal{L}_0+ \mathcal{L}_\textrm{int}\,\) as in the untransformed case and abbreviate the additional term, which is due to the exponential \(e^{v{\bar{v}}}\) which in turn comes from the initial state \(\psi _0\) which was chosen to be a product of coherent states, as

$$\begin{aligned} \mathcal{L}_{\psi _0}:= & {} 2u \,{\sum }_j v_{j}{\bar{v}}_{j}\, { \Bigl \{\, v_j \, {\partial \over \partial v_j}\,-\,{\bar{v}}_j \, {\partial \over \partial \bar{v}_j} \,\Bigr \} } \end{aligned}$$
(3.30)

Then the solution of the PDE (3.27) can be written as

$$\begin{aligned} F_t= & {} e^{-it(\mathcal{L}+\mathcal{L}_{\psi _0})}\,F_0 \end{aligned}$$
(3.31)

For the density matrix elements, we now have to calculate the expectation (3.20) instead of (3.1) and hence the initial condition \(F_0\) is simply given by

$$\begin{aligned} F_0(\,\{v_j\},\{\bar{v}_j\}\,)= & {} v_i\,{\bar{v}}_j \end{aligned}$$
(3.32)

instead of the \(\,v_i\,{\bar{v}}_j\; e^{\,v{\bar{v}}\,-\,|{\lambda }|^2}\,\) which we had in the untransformed case. Thus, the time evolved density matrix elements are obtained through

$$\begin{aligned} (\,\psi _t\,,a_i^+a_j \,\psi _t\,)_\mathcal{F}= & {} e^{-it(\mathcal{L}_0+ \mathcal{L}_\textrm{int}+\mathcal{L}_{\psi _0})}\,\{\, v_i\,{\bar{v}}_j \,\}\, \bigr |_{v={\lambda },{\bar{v}}={\bar{\lambda }}} \end{aligned}$$
(3.33)

Again, the solution of the PDE (3.27) is not needed in the whole \((v,{\bar{v}})\) - space, but only at one specific point \((v,\bar{v})=({\lambda },\bar{\lambda })\).

3.3 The PDE Version of the Girsanov Transformation Formula

By comparison of (3.18) and (3.32), apparently there has to be the identity

$$\begin{aligned} e^{-it(\mathcal{L}_0+\mathcal{L}_\textrm{int})}\,\bigl \{\, f(v,{\bar{v}})\; e^{\,v{\bar{v}}\,-\,|{\lambda }|^2} \,\bigr \}\, \bigr |_{v={\lambda },{\bar{v}}={\bar{\lambda }}}= & {} e^{-it(\mathcal{L}_0+\mathcal{L}_\textrm{int}+\mathcal{L}_{\psi _0})}\,\{\, f(v,{\bar{v}}) \,\}\, \bigr |_{v={\lambda },{\bar{v}}={\bar{\lambda }}} \end{aligned}$$
(3.34)

for arbitrary functions \(f(v,{\bar{v}})\). This identity has the following generalization which can be obtained by redoing the calculations of chapter 1 and 2 for an arbitrary initial state, not necessarily for a coherent state: Let \(P=P(v{\bar{v}})\) be some (usually positive when restricted to real values, since it come from a \(\Vert \psi _t\Vert ^2\) ) arbitrary function of one real or complex variable. Then

$$\begin{aligned} e^{-it(\mathcal{L}_0+\mathcal{L}_\textrm{int})}\,\bigl \{\, f(v,{\bar{v}})\; {P(v{\bar{v}})\over P({\lambda }{\bar{\lambda }})} \,\bigr \}\, \bigr |_{v={\lambda },{\bar{v}}={\bar{\lambda }}}= & {} e^{-it(\mathcal{L}_0+\mathcal{L}_\textrm{int}+\mathcal{L}_{P})}\,\{\, f(v,{\bar{v}}) \,\}\, \bigr |_{v={\lambda },{\bar{v}}={\bar{\lambda }}} \end{aligned}$$
(3.35)

or equivalently

$$\begin{aligned} e^{-it(\mathcal{L}_0+\mathcal{L}_\textrm{int})}\,\bigl \{\, P(v{\bar{v}})\,f(v,{\bar{v}}) \,\bigr \}= & {} P(v{\bar{v}})\;\times \; e^{-it(\mathcal{L}_0+\mathcal{L}_\textrm{int}+\mathcal{L}_{P})}\,\{\, f(v,{\bar{v}}) \,\} \end{aligned}$$
(3.36)

with

$$\begin{aligned} \mathcal{L}_{P}:= & {} 2u \,{ { P'(v{\bar{v}})\over P(v{\bar{v}})}} \;{\sum }_j v_{j}{\bar{v}}_{j}\, { \Bigl \{\, v_j \, {\partial \over \partial v_j}\,-\,{\bar{v}}_j \, {\partial \over \partial \bar{v}_j} \,\Bigr \} } \end{aligned}$$
(3.37)

and \(P'(x)=dP/dx\).

Since (3.35) is completely independent of any stochastics, just some algebraic statement concerning derivatives, let us also give an independent proof, thereby confirming the validity of the stochastic formalismFootnote 1 which has been used so far: First, one calculates that for any \(\,P=P(v{\bar{v}})=P\bigl (\Sigma _j v_j{\bar{v}}_j)\,\), one has, using \(\varepsilon _{ij}=\varepsilon _{ji}\) for the \(\mathcal{L}_0\) equation,

$$\begin{aligned} \mathcal{L}_0P= & {} 0 \end{aligned}$$
(3.38)
$$\begin{aligned} \mathcal{L}_\textrm{int}P= & {} 0 \end{aligned}$$
(3.39)

Since \(\mathcal{L}_0\) is a first order operator, one has for arbitrary \(f=f(\,\{v_j\},\{\bar{v}_j\} \,)\)

$$\begin{aligned} \mathcal{L}_0(Pf)= & {} \mathcal{L}_0(P) \times f\;+\; P\times \mathcal{L}_0(f) \;\;=\;\; P\times \mathcal{L}_0(f) \end{aligned}$$
(3.40)

Furthermore,

$$\begin{aligned} \mathcal{L}_\textrm{int}(Pf)= & {} u\;{\sum }_j \Bigl \{\; v_j^2\, {\partial ^2 \over \partial v_j^2} \,-\,\bar{v}_j^2\, {\partial ^2 \over \partial \bar{v}_j^2} \;\Bigr \} (Pf) \nonumber \\= & {} \mathcal{L}_\textrm{int}(P)\times f\;+\; P\times \mathcal{L}_\textrm{int}(f)\;+\; 2u\;{\sum }_j \Bigl \{\; v_j^2\, {\partial P \over \partial v_j}\,{\partial f \over \partial v_j} \,-\,\bar{v}_j^2\, {\partial P \over \partial \bar{v}_j}\,{\partial f \over \partial \bar{v}_j} \;\Bigr \} \nonumber \\= & {} P\times \mathcal{L}_\textrm{int}(f)\;+\; 2u\;{\sum }_j \Bigl \{\; v_j^2\, P'\,{\bar{v}}_j\,{\partial f \over \partial v_j} \,-\,\bar{v}_j^2\, P'\,v_j\,{\partial f \over \partial \bar{v}_j} \;\Bigr \} \nonumber \\= & {} P\times \mathcal{L}_\textrm{int}(f)\;+\; 2uP'\;{\sum }_j v_j{\bar{v}}_j\,\Bigl \{\; v_j\,{\partial f \over \partial v_j} \,-\,\bar{v}_j\,{\partial f \over \partial \bar{v}_j} \;\Bigr \} \nonumber \\= & {} P\times \bigl [\,\mathcal{L}_\textrm{int}(f) \,+\,\mathcal{L}_P(f) \, \bigr ] \end{aligned}$$
(3.41)

Thus,

$$\begin{aligned} (\mathcal{L}_0+\mathcal{L}_\textrm{int})(Pf)= & {} P\;\times \;(\mathcal{L}_0+\mathcal{L}_\textrm{int}+\mathcal{L}_P)(f) \end{aligned}$$
(3.42)

and by induction

$$\begin{aligned} (\mathcal{L}_0+\mathcal{L}_\textrm{int})^{n+1}(Pf)= & {} (\mathcal{L}_0+\mathcal{L}_\textrm{int})\,\bigl \{\, (\mathcal{L}_0+\mathcal{L}_\textrm{int})^{n}(Pf) \,\bigr \}\nonumber \\= & {} (\mathcal{L}_0+\mathcal{L}_\textrm{int})\,\bigl \{\, P\,\times \, (\mathcal{L}_0+\mathcal{L}_\textrm{int}+\mathcal{L}_P)^{n}(f) \,\bigr \}\nonumber \\&\buildrel \text {(3.41)}\over =&P\;\times \; (\mathcal{L}_0+\mathcal{L}_\textrm{int}+\mathcal{L}_P)\,\Bigl \{\, (\mathcal{L}_0+\mathcal{L}_\textrm{int}+\mathcal{L}_P)^{n}(f) \, \Bigr \}\nonumber \\= & {} P\;\times \; (\mathcal{L}_0+\mathcal{L}_\textrm{int}+\mathcal{L}_P)^{n+1}(f) \end{aligned}$$
(3.43)

This proves (3.35). Let us summarize the results of this chapter in the following

Theorem 6

For an arbitrary real symmetric hopping matrix \(\varepsilon =(\varepsilon _{ij})=(\varepsilon _{ji})\), define the differential operators

$$\begin{aligned} \mathcal{L}_0:= & {} {\sum }_j \Bigl \{\; { (\varepsilon v)_j\, {\partial \over \partial v_j} \,-\, (\varepsilon \bar{v})_j\, {\partial \over \partial \bar{v}_j} } \;\Bigr \} \;\;=\;\;{\sum }_{i,j} { \,\varepsilon _{ij}\Bigl \{\; v_i\, {\partial \over \partial v_j} \,-\, \bar{v}_i\, {\partial \over \partial \bar{v}_j} } \;\Bigr \} \end{aligned}$$
(3.44)
$$\begin{aligned} \mathcal{L}_\textrm{int}:= & {} u\;{\sum }_j \Bigl \{\; v_j^2\, {\partial ^2 \over \partial v_j^2} \,-\,\bar{v}_j^2\, {\partial ^2 \over \partial \bar{v}_j^2} \;\Bigr \} \end{aligned}$$
(3.45)

Furthermore, for some arbitrary function of one variable \(\,P=P(v{\bar{v}})=P\bigl (\Sigma _j v_j{\bar{v}}_j\bigr )\),  put

$$\begin{aligned} \mathcal{L}_{P}:= & {} 2u \,{ { P'(v{\bar{v}})\over P(v{\bar{v}})}} \;{\sum }_j v_{j}{\bar{v}}_{j}\, { \Bigl \{\, v_j \, {\partial \over \partial v_j}\,-\,{\bar{v}}_j \, {\partial \over \partial \bar{v}_j} \,\Bigr \} } \end{aligned}$$
(3.46)

Then:

  • a) The density matrix elements of Theorem 2 can be calculated through the following formula, this is the untransformed case before Girsanov transformation:

    $$\begin{aligned} (\,\psi _t\,,a_i^+a_j \,\psi _t\,)_\mathcal{F}= & {} e^{-it(\mathcal{L}_0+\mathcal{L}_\textrm{int})}\,\bigl \{\, v_i\,{\bar{v}}_j\; e^{\,v{\bar{v}}\,-\,|{\lambda }|^2} \,\bigr \}\, \bigr |_{v={\lambda },{\bar{v}}={\bar{\lambda }}} \end{aligned}$$
    (3.47)
  • b) The density matrix elements of Theorem 2 can be calculated through the following equivalent formula, obtained from the SDE representation (2.59),(2.60) after Girsanov transformation:

    $$\begin{aligned} (\,\psi _t\,,a_i^+a_j \,\psi _t\,)_\mathcal{F}= & {} e^{-it(\mathcal{L}_0+\mathcal{L}_\textrm{int}+\mathcal{L}_P)}\,\bigl \{\, v_i\,{\bar{v}}_j \,\bigr \}\, \bigr |_{v={\lambda },{\bar{v}}={\bar{\lambda }}} \end{aligned}$$
    (3.48)

    with \(P(x)=e^{x}\).

  • c) For an arbitrary function \(P=P(x)\) of one variable, there is the following general identity:

    $$\begin{aligned} e^{-it(\mathcal{L}_0+\mathcal{L}_\textrm{int})}\,\bigl \{\, P(v{\bar{v}})\,f(v,{\bar{v}}) \,\bigr \}= & {} P(v{\bar{v}})\;\times \; e^{-it(\mathcal{L}_0+\mathcal{L}_\textrm{int}+\mathcal{L}_{P})}\,\{\, f(v,{\bar{v}}) \,\} \end{aligned}$$
    (4.1)

    with \(f=f(v,{\bar{v}})=f(\,\{v_j\},\{\bar{v}_j\} \,)\,\) being an arbitrary function of \(2{|\Gamma |}\) variables.

4 An Explicit Solvable Test Case: The 0D Bose-Hubbard Model

Since stochastic calculus with Fresnel Brownian motions instead of standard Brownian motions is not part of rigorous textbook mathematics, let’s make an additional check of the formalism by applying it to an explicit solvable test case, the 0D Bose-Hubbard model. The term 0D Bose-Hubbard model we borrowed from the paper [8] of Ray, Ostmann, Simon, Grossmann and Strunz, where the model also had to serve as a test example. Besides of just being a test case, the purpose of this chapter is also to give some intuition for the approximation which will be used in Section 5.3 to take the diffusive part of the SDE system into account.

In the Bargmann-Segal representation (1.4), the Hamiltonian of the 0D Bose-Hubbard model is simply

$$\begin{aligned} h\;\;=\;\; h_0\,+\,h_\textrm{int}:= & {} \varepsilon \,z {d\over dz} \;+\; u\, z^2 {d^2\over dz^2} \end{aligned}$$
(4.2)

We choose the initial state

$$\begin{aligned} \psi _0(z):= & {} e^{\lambda z} \;e^{-|\lambda |^2/2} \end{aligned}$$
(4.3)

with \(\lambda \in \mathbb C\) and consider the time evolution

$$\begin{aligned} \psi _t:= & {} e^{-ith}\psi _0 \end{aligned}$$
(4.4)

We want to calculate the quantity, with \(\,a=d/dz\) , \(\,a^+=z\,\),

$$\begin{aligned} (\psi _t,a\,\psi _t)_\mathcal{F}= & {} (a^+\psi _t,\psi _t)_\mathcal{F}\;\;=\;\; \int _{\mathbb C} d\mu (z)\;z\;|\psi _t(z)|^2 \nonumber \\= & {} \int _{\mathbb R^2} {d\textrm{Re}z\,d\textrm{Im}z\over \pi }\; e^{-|z|^2} \; z\;|\psi _t(z)|^2 \end{aligned}$$
(4.5)

Since

$$\begin{aligned} h\, z^n= & {} \bigl \{\varepsilon n +un(n-1)\bigr \}\,z^n \;\;=:\;\; h_n\,z^n \end{aligned}$$
(4.6)

we have

$$\begin{aligned} \psi _t(z)= & {} \sum _{n=0}^\infty {\lambda ^n\over n!}\; e^{-ith_n}\,z^n \;e^{-|\lambda |^2/2} \end{aligned}$$
(4.7)

such that

$$\begin{aligned} (\psi _t,a\,\psi _t)_\mathcal{F}= & {} \sum _{n,m=0}^\infty e^{-it(h_n-h_m)}\; {\lambda ^n\bar{\lambda }^m\over n!\,m!}\; \int _{\mathbb R^2} {d^2z\over \pi }\; e^{-|z|^2} \; {z}^{n+1} {\bar{z}}^{m}\;e^{-|\lambda |^2} \nonumber \\= & {} \sum _{n=0}^\infty e^{-it(h_{n}-h_{n+1})}\; {\lambda ^{n}{\bar{\lambda }}^{n+1}\over n!\,(n+1)!}\; \int _{\mathbb R^2} {d^2z\over \pi }\; e^{-|z|^2} \; |z|^{2(n+1)}\;e^{-|\lambda |^2} \nonumber \\= & {} \sum _{n=0}^\infty e^{+it(h_{n+1}-h_{n})}\; {\lambda ^{n}{\bar{\lambda }}^{n+1}\over n!}\;\;e^{-|\lambda |^2} \end{aligned}$$
(4.8)

Since

$$\begin{aligned} h_{n+1}-h_n= & {} \varepsilon -u+u(2n+1)\;\;=\;\;\varepsilon +2u\,n \end{aligned}$$
(4.9)

we end up with

$$\begin{aligned} (\psi _t,a\,\psi _t)_\mathcal{F}= & {} \bar{\lambda }\, \sum _{n=0}^\infty e^{+it(\varepsilon +2u\,n )}\; { {(\lambda {\bar{\lambda }})^n\over n!} }\;e^{-|\lambda |^2} \nonumber \\= & {} \bar{\lambda }\; e^{i\varepsilon t}\; \exp \bigl \{\, -\,(1-e^{2iut})\,|\lambda |^2\,\bigr \} \end{aligned}$$
(4.10)

This quantity has already collapse and revivals, if we plot \(\,\textrm{Re}(\psi _t,a\,\psi _t)_\mathcal{F}\,\) for \(\,\varepsilon =2\,\), \(\,u=g/N\,\) with \(\,g=0.5\,\) and \(\,N=|{\lambda }|^2=20\,\), we get

figure a

Let’s apply Theorem 2. We get the following representation:

$$\begin{aligned} (\,\psi _t\,,a \,\psi _t\,)_\mathcal{F}= & {} \textsf{E}_x\bar{\textsf{E}}_y\bigl [ \;{\bar{v}}_{y,t}\; e^{v_{x,t}{\bar{v}}_{y,t}} \;\bigr ] \;e^{-|{\lambda }|^2} \end{aligned}$$
(4.11)

where the \(v_t,{\bar{v}}_t\in \mathbb C\) are given by the SDEs

$$\begin{aligned} dv_t= & {} \;-\;i\, dt\,\varepsilon \, v_t \;-\;i\,\sqrt{2u}\,v_t\,dx_t \nonumber \\ d\bar{v}_t= & {} \;+\;i\, dt\,\varepsilon \,\bar{v}_t \;+\;i\,\sqrt{2u}\,\bar{v}_t\,dy_t \end{aligned}$$
(4.12)

with initial conditions \(\,v_{0}={\lambda },\,{\bar{v}}_{0}=\bar{\lambda }\,\). If we would have Wiener measure instead of Fresnel measure, this would be a geometric Brownian motion. Here we have Fresnel measure with calculation rules

$$\begin{aligned} (dx_t)^2= & {} +i\,dt \nonumber \\ (dy_t)^2= & {} -i\,dt \end{aligned}$$
(4.13)

and obtain the solutions

$$\begin{aligned} v_t= & {} v_0\, e^{-i(\varepsilon -u)t\,-\,i\sqrt{2u}\,x_t } \nonumber \\ {\bar{v}}_t= & {} {\bar{v}}_0\, e^{+i(\varepsilon -u)t\,+\,i\sqrt{2u}\,y_t } \end{aligned}$$
(4.14)

Namely, the Ito formula applied to \(\,v_t=v_0\, e^{-i(\varepsilon -u)t\,-\,i\sqrt{2u}\,x_t }= v(x_t,t)\,\) gives

$$\begin{aligned} dv_t= & {} {\partial v \over \partial x} \, dx_t \;+\; {1\over 2}\,{\partial ^2 v \over \partial x^2}\,(dx_t)^2 \;+\; {\partial v \over \partial t} \, dt \nonumber \\= & {} \,-\,i\sqrt{2u}\,v_t\,dx_t \;+\; {1\over 2}\,(-i\sqrt{2u})^2\,v_t\,(dx_t)^2 \;-\;i\,(\varepsilon -u)\,v_t\, dt \nonumber \\&\buildrel \text {(4.12)}\over =&\,-\,i\sqrt{2u}\,v_t\,dx_t \;+\; {1\over 2}\,(-i\sqrt{2u})^2\,v_t\;(+i)\,dt \;-\;i\,(\varepsilon -u)\,v_t\, dt \nonumber \\= & {} \,-\,i\sqrt{2u}\,v_t\,dx_t \;-\;i\,\varepsilon \,dt\, v_t \end{aligned}$$
(4.15)

For some integrand which depends on \(x_t\) only and not on earlier \(\{x_{t'}\}_{t'<t}\,\), Fresnel or Wiener expectations reduce to 1-dimensional integrals (Appendix A.4 has a general formula in (A.84)). We have

$$\begin{aligned} \textsf{E}_x\bar{\textsf{E}}_y\bigl [ \;f(x_t,y_t)\;\bigr ]= & {} \int _{{\mathbb R}^2} f(x_t,y_t)\; e^{\,i{x_t^2-y_t^2\over 2t}} \,{dx_t\,dy_t\over 2\pi t} \nonumber \\= & {} \int _{{\mathbb R}^2} \,f(\,\sqrt{t}\,x\,,\,\sqrt{t}\,y\,)\; e^{\,i{x^2-y^2\over 2}} \,{dx\,dy\over 2\pi } \end{aligned}$$
(4.16)

Now we substitute

$$\begin{aligned} \xi \;\;=\;\;{x-y\over \sqrt{2}}\;,\;\;\;\; \eta \;\;=\;\;{x+y\over \sqrt{2}} \end{aligned}$$
(4.17)

which gives

$$\begin{aligned} e^{i{x^2-y^2\over 2}} \,{dx\,dy\over 2\pi }= & {} e^{i\xi \eta } \;{d\xi \,d\eta \over 2\pi } \end{aligned}$$
(4.18)

The quantities in the integrand in (4.15) are given by (4.10),

$$\begin{aligned} \;e^{v_{x,t}{\bar{v}}_{y,t}}= & {} \exp \bigl \{\, |{\lambda }|^2\,e^{\,-\,i\sqrt{2u}\,(x_t-y_t)\,} \,\bigr \} \;\;=\;\;\exp \bigl \{\, |{\lambda }|^2\,e^{\,-\,i\sqrt{4ut}\,\xi \,} \,\bigr \} \end{aligned}$$
(4.19)

and

$$\begin{aligned} \;{\bar{v}}_{y,t}\; e^{v_{x,t}{\bar{v}}_{y,t}}= & {} \bar{\lambda }\; e^{+i(\varepsilon -u)t\,+\,i\sqrt{2u}\,y_t } \;\times \; \exp \bigl \{\, |{\lambda }|^2\,e^{\,-\,i\sqrt{4ut}\,\xi \,} \,\bigr \} \nonumber \\= & {} \bar{\lambda }\; e^{+i(\varepsilon -u)t} \; e^{\,+\,i\sqrt{ut}\,(\eta -\xi ) } \;\times \; \exp \bigl \{\, |{\lambda }|^2\,e^{\,-\,i\sqrt{4ut}\,\xi \,} \,\bigr \} \end{aligned}$$
(4.20)

Thus we have to evaluate

$$\begin{aligned} (\,\psi _t\,,a \,\psi _t\,)_\mathcal{F}= & {} \textsf{E}_x\bar{\textsf{E}}_y\bigl [ \;{\bar{v}}_{y,t}\; e^{v_{x,t}{\bar{v}}_{y,t}} \;\bigr ] \;e^{-|{\lambda }|^2} \nonumber \\= & {} \int _{{\mathbb R}^2} \bar{\lambda }\, e^{+i(\varepsilon -u)t} \; e^{\,+\,i\sqrt{ut}\,(\eta -\xi ) } \,\times \, \exp \bigl \{\, |{\lambda }|^2\,(e^{\,-\,i\sqrt{4ut}\,\xi \,}-1) \,\bigr \} \; e^{i\xi \eta } \;{d\xi \,d\eta \over 2\pi } \nonumber \\= & {} \bar{\lambda }\; e^{+i(\varepsilon -u)t} \int _{{\mathbb R}} d\xi \; e^{\,-\,i\sqrt{ut}\,\xi } \, \int _{{\mathbb R}} e^{\,+\,i(\xi +\sqrt{ut})\,\eta } \;{d\eta \over 2\pi } \; \times \; \exp \bigl \{\, |{\lambda }|^2\,(e^{\,-\,i\sqrt{4ut}\,\xi \,}-1) \,\bigr \} \nonumber \\= & {} \bar{\lambda }\; e^{+i(\varepsilon -u)t} \int _{{\mathbb R}} d\xi \; e^{\,-\,i\sqrt{ut}\,\xi } \;\delta (\xi +\sqrt{ut}) \; \times \; \exp \bigl \{\, |{\lambda }|^2\,(e^{\,-\,i\sqrt{4ut}\,\xi \,}-1) \,\bigr \} \nonumber \\= & {} \bar{\lambda }\; e^{+i(\varepsilon -u)t} \; e^{\,+\,iut } \; \times \; \exp \bigl \{\, |{\lambda }|^2\,(e^{\,+\,i\,2ut\,}-1) \,\bigr \} \nonumber \\= & {} \bar{\lambda }\; e^{+i\varepsilon t} \; \times \; \exp \bigl \{\, |{\lambda }|^2\,(e^{\,+\,i\,2ut\,}-1) \,\bigr \} \end{aligned}$$
(4.21)

and this coincides with the exact result (4.9). Now let’s apply Theorem 4. After Girsanov transformation, we have the following representation (we omit the tilde on the transformed Fresnel BMs in the following):

$$\begin{aligned} (\,\psi _t\,,a \,\psi _t\,)_\mathcal{F}= & {} \textsf{E}_x\bar{\textsf{E}}_y\bigl [ \;{\bar{v}}_{t} \;\bigr ] \end{aligned}$$
(4.22)

with \(v,{\bar{v}}\) given by

$$\begin{aligned} dv_t= & {} \;-\;i\, dt\,\varepsilon \, v_t \;-\;i\,2u\,dt\,v_t{\bar{v}}_t\,v_t \;-\;i\,\sqrt{2u}\,v_t\,dx_t \nonumber \\ d\bar{v}_t= & {} \;+\;i\, dt\,\varepsilon \,\bar{v}_t \;+\;i\,2u\,dt\,v_t{\bar{v}}_t\,{\bar{v}}_t \;+\;i\,\sqrt{2u}\,\bar{v}_t\,dy_t \end{aligned}$$
(4.23)

This SDE system can still be solved in closed form. First we calculate

$$\begin{aligned} d(v{\bar{v}})= & {} dv\,{\bar{v}}\;+\;v\,d{\bar{v}}\;+\; dv\,d{\bar{v}}\nonumber \\= & {} \;-\;i\, dt\,\varepsilon \, v{\bar{v}}\;-\;i\,2u\,dt\,v{\bar{v}}\,v{\bar{v}}\;-\;i\,\sqrt{2u}\,v{\bar{v}}\,dx_t \nonumber \\{} & {} \;+\;i\, dt\,\varepsilon \,v\bar{v} \;+\;i\,2u\,dt\,v{\bar{v}}\,v{\bar{v}}\;+\;i\,\sqrt{2u}\,v\bar{v}\,dy_t \;\;+\;\;0 \nonumber \\= & {} \;-\;i\,\sqrt{2u}\,v{\bar{v}}\,(dx_t-dy_t) \;\;=\;\; \;-\;i\,\sqrt{4u}\,v{\bar{v}}\, d\xi _t \end{aligned}$$
(4.24)

Since

$$\begin{aligned} (d\xi _t)^2\;\;\!=\!\;\; \bigl (\,{dx_t-dy_t\over \sqrt{2}}\,\bigr )^2 \;\;\!=\!\;\; { (dx_t)^2 \,-\,2dx_tdy_t \,\!+\!\,(dy_t)^2 \over 2 } \;\;=\;\; { +i\,dt \,-\,0\,-\,i\,dt \over 2} \;\;=\;\; 0 \ \ \ \ \ \ \ \end{aligned}$$
(4.25)

the solution to (4.23) is

$$\begin{aligned} (v{\bar{v}})_t= & {} (v{\bar{v}})_0\; e^{-i\sqrt{4u}\,\xi _t} \;\;=\;\;|{\lambda }|^2\; e^{-i\sqrt{4u}\,\xi _t} \end{aligned}$$
(4.26)

Thus, the equation for \({\bar{v}}\) becomes

$$\begin{aligned} d{\bar{v}}= & {} \Bigl [\;+\;i\,\bigl (\,\varepsilon \,+\,2u|{\lambda }|^2\, e^{-i\sqrt{4u}\,\xi _t} \,\bigr )\, dt \;+\;i\,\sqrt{2u}\,dy_t \;\Bigr ]\, {\bar{v}}\end{aligned}$$
(4.27)

This is a geometric Fresnel BM with a time dependent and stochastic drift. To solve it, we have to take into account that \(\,(dy_t)^2=-i\,dt\,\) and obtain

$$\begin{aligned} {\bar{v}}_t= & {} {\bar{v}}_0\; e^{\,+\,i\int _0^t\bigl (\,\varepsilon \,+\,2u|{\lambda }|^2\, e^{-i\sqrt{4u}\,\xi _s} \,\bigr )\, ds \,+\,i\sqrt{2u}\,y_t\,-\,iut } \nonumber \\= & {} \bar{\lambda }\;e^{+i(\varepsilon -u)t}\;e^{\,+\,i\,2u|{\lambda }|^2\int _0^t e^{-i\sqrt{4u}\,\xi _s}\, ds \;+\;i\sqrt{2u}\,y_t\, } \nonumber \\= & {} \bar{\lambda }\;e^{+i(\varepsilon -u)t}\;e^{\,+\,i\,2u|{\lambda }|^2\int _0^t e^{-i\sqrt{4u}\,\xi _s}\, ds \;+\;i\sqrt{u}\,(\eta _t-\xi _t)\, } \end{aligned}$$
(4.28)

Since \({\bar{v}}_t\) is not just a function of the Fresnel BMs at time t, but it depends also through the ds-integral in the exponent on the Fresnel BMs \(\xi _s\) at earlier times \(s<t\), we can’t no longer make a large step evaluation of the expectation value through a one or two dimensional integral, but we have to write down the full small step path integral. Recall the notations

$$\begin{aligned} x_{t_k}= & {} \sqrt{dt}\,{\sum }_{\ell =1}^k \phi _\ell \nonumber \\ y_{t_k}= & {} \sqrt{dt}\,{\sum }_{\ell =1}^k \theta _\ell \end{aligned}$$
(4.29)

and the Fresnel measure

$$\begin{aligned} dF\,d\bar{F}= & {} \mathop \Pi _{\ell =1}^k e^{\,i\,{\phi _\ell ^2-\theta _\ell ^2\over 2}}\; {d\phi _\ell \,d\theta _\ell \over 2\pi } \end{aligned}$$
(4.30)

Let us write

$$\begin{aligned} \xi _{t_k}\;\;=\;\; {x_{t_k}-y_{t_k}\over \sqrt{2}} \;\;=\;\;\sqrt{dt}\,{\sum }_{\ell =1}^k {\phi _\ell -\theta _\ell \over \sqrt{2}}=: & {} \sqrt{dt}\,{\sum }_{\ell =1}^k \alpha _\ell \nonumber \\ \eta _{t_k}\;\;=\;\; {x_{t_k}+y_{t_k}\over \sqrt{2}} \;\;=\;\;\sqrt{dt}\,{\sum }_{\ell =1}^k {\phi _\ell +\theta _\ell \over \sqrt{2}}=: & {} \sqrt{dt}\,{\sum }_{\ell =1}^k \beta _\ell \end{aligned}$$
(4.31)

such that the Fresnel measure becomes

$$\begin{aligned} dF\,d\bar{F}= & {} \mathop \Pi _{\ell =1}^k e^{\,i\,\alpha _\ell \beta _\ell }\; {d\alpha _\ell \,d\beta _\ell \over 2\pi } \end{aligned}$$
(4.32)

Then in discrete time \(t=t_k=kdt\,\) the \({\bar{v}}\) is given by

$$\begin{aligned} {\bar{v}}_{t_k}= & {} \bar{\lambda }\;e^{+i(\varepsilon -u)t_k}\; \exp \Bigl \{\,+\,i\,2u|{\lambda }|^2{\sum }_{\ell =1}^k e^{-i\sqrt{4u}\,\xi _{t_\ell }}\, dt \;+\;i\sqrt{udt}\,{\sum }_{\ell =1}^k(\beta _\ell -\alpha _\ell )\, \Bigr \} \ \ \ \ \ \ \ \end{aligned}$$
(4.33)

We have to calculate

$$\begin{aligned} (\,\psi _t\,,a \,\psi _t\,)_\mathcal{F}= & {} \textsf{E}_x\bar{\textsf{E}}_y\bigl [ \;{\bar{v}}_{t_k} \;\bigr ] \;\;=\;\; { \int _{{\mathbb R}^{2k}} \;{\bar{v}}_{t_k}(\alpha ,\beta ) }\; \mathop \Pi _{\ell =1}^k e^{\,i\,\alpha _\ell \beta _\ell }\; {d\alpha _\ell \,d\beta _\ell \over 2\pi } \end{aligned}$$
(4.34)

The \(\beta \)-integrals produce \(\delta \)-functions,

$$\begin{aligned} { \int _{{\mathbb R}^{k}} \exp \bigl \{\;+\;i\sqrt{udt}\,{\sum }_{\ell =1}^k \beta _\ell \, \bigr \} }\; \mathop \Pi _{\ell =1}^k e^{\,i\,\alpha _\ell \beta _\ell }\; {d\beta _\ell \over 2\pi }= & {} \mathop \Pi _{\ell =1}^k \,\delta (\sqrt{udt}+\alpha _\ell ) \end{aligned}$$
(4.35)

Thus we have \(\,\alpha _\ell =-\sqrt{udt}\,\) for all \(\ell \) and the \(\xi _{t_\ell }\) in (4.32) becomes

$$\begin{aligned} \xi _{t_\ell }= & {} \sqrt{dt}\,{\sum }_{m=1}^\ell \alpha _m \;\;=\;\; -\,\sqrt{u}\,dt\,\ell \;\;=\;\; -\, \sqrt{u}\,t_\ell \end{aligned}$$
(4.36)

We end up with

$$\begin{aligned} \textsf{E}_x\bar{\textsf{E}}_y\bigl [ \;{\bar{v}}_{t_k} \;\bigr ]= & {} \bar{\lambda }\;e^{+i(\varepsilon -u)t_k}\; \exp \Bigl \{\,+\,i\,2u|{\lambda }|^2{\sum }_{\ell =1}^k e^{+i\sqrt{4u}\,\sqrt{u}\,t_\ell }\, dt \;+\;i\sqrt{udt}\,{\sum }_{\ell =1}^k\,\sqrt{udt} \; \Bigr \} \nonumber \\= & {} \bar{\lambda }\;e^{+i(\varepsilon -u)t_k}\; \exp \Bigl \{\,+\,i\,2u|{\lambda }|^2{\sum }_{\ell =1}^k e^{+i\,2u\,t_\ell }\, dt \;+\;i\,udt\, k \; \Bigr \} \nonumber \\&\buildrel dt\rightarrow 0 \over =&\bar{\lambda }\;e^{+i \varepsilon t}\; \exp \Bigl \{\,+\,i\,2u|{\lambda }|^2\int _0^t e^{+i\,2u\,s}\, ds \; \Bigr \} \nonumber \\= & {} \bar{\lambda }\;e^{+i \varepsilon t}\; \exp \Bigl \{\;|{\lambda }|^2( e^{+i\,2u\,t}-1) \; \Bigr \} \end{aligned}$$
(4.37)

and this again coincides with the original result (4.9).

Finally, let us check the PDE representations of chapter 3 which are summarized in Theorem 6. There we have the operators \(\mathcal{L}_0\), \(\mathcal{L}_\textrm{int}\) and \(\mathcal{L}_P\) with \(P(x)=e^x\) which for the current 0D case reduce to

$$\begin{aligned} \mathcal{L}_0= & {} \varepsilon \, \bigl \{\, v\, {\partial \over \partial v} \,-\, \bar{v}\, {\partial \over \partial \bar{v}} \,\bigr \} \nonumber \\ \mathcal{L}_\textrm{int}= & {} u\,\bigl \{\, v^2\, {\partial ^2 \over \partial v^2} \,-\,\bar{v}^2\, {\partial ^2 \over \partial \bar{v}^2} \,\bigr \} \\ \mathcal{L}_{P}= & {} 2u \,v{\bar{v}}\, \bigl \{\, v \, {\partial \over \partial v}\,-\,{\bar{v}}\, {\partial \over \partial \bar{v}} \,\bigr \} \nonumber \end{aligned}$$
(4.38)

Theorem 6 makes the following statements:

  1. a)

    Untransformed Representation:

    $$\begin{aligned} (\,\psi _t\,,a \,\psi _t\,)_\mathcal{F}&\buildrel !\over =&e^{-it(\mathcal{L}_0+\mathcal{L}_\textrm{int})}\,\bigl \{\,{\bar{v}}\; e^{\,v{\bar{v}}\,-\,|{\lambda }|^2} \,\bigr \}\, \bigr |_{v={\lambda },{\bar{v}}={\bar{\lambda }}} \end{aligned}$$
    (4.39)
  2. b)

    Transformed Representation:

    $$\begin{aligned} (\,\psi _t\,,a \,\psi _t\,)_\mathcal{F}&\buildrel !\over =&e^{-it(\mathcal{L}_0+\mathcal{L}_\textrm{int}+\mathcal{L}_P)}\,\bigl \{\,{\bar{v}}\,\bigr \}\, \bigr |_{v={\lambda },{\bar{v}}={\bar{\lambda }}} \end{aligned}$$
    (4.40)

In case (b), we have to show that the function

$$\begin{aligned} F_t(v,{\bar{v}}):= & {} {\bar{v}}\;e^{+i \varepsilon t}\, \exp \bigl \{\,v{\bar{v}}( e^{+i\,2u\,t}-1) \, \bigr \} \end{aligned}$$
(4.41)

this is the original result (4.9) with the \(({\lambda },{\bar{\lambda }})\) replaced by \((v_0,{\bar{v}}_0)\rightarrow (v,{\bar{v}})\), is a solution of the PDE

$$\begin{aligned} i\,{\partial \over \partial t}\, F_t= & {} (\mathcal{L}_0+\mathcal{L}_\textrm{int}+\mathcal{L}_P)\,F_t \end{aligned}$$
(4.42)

with initial condition \(F_0={\bar{v}}\,\). The initial condition is obvious, so let’s check the derivatives. We have

$$\begin{aligned} i\,{\partial \over \partial t}\, F_t= & {} \,-\,\varepsilon \,F_t\;-\;2u\,e^{+i2ut}v{\bar{v}}\, F_t \end{aligned}$$
(4.43)

and one calculates

$$\begin{aligned} \mathcal{L}_0\,F_t= & {} \,-\,\varepsilon \,F_t \nonumber \\ \mathcal{L}_\textrm{int}\,F_t= & {} \,-\,2u\,(e^{+i2ut}-1)v{\bar{v}}\, F_t \\ \mathcal{L}_P\,F_t= & {} \,-\,2u\,v{\bar{v}}\, F_t \nonumber \end{aligned}$$
(4.44)

which validates (4.41) or (4.39).

In case (a), we have to show that the function

$$\begin{aligned} G_t(v,{\bar{v}}):= & {} F_t(v,{\bar{v}})\;e^{\,+\,v{\bar{v}}}\;\;=\;\; {\bar{v}}\;e^{+i \varepsilon t}\, \exp \bigl \{\,v{\bar{v}}\, e^{+i\,2u\,t} \, \bigr \} \end{aligned}$$
(4.45)

is a solution of the PDE

$$\begin{aligned} i\,{\partial \over \partial t}\, G_t= & {} (\mathcal{L}_0+\mathcal{L}_\textrm{int})\,G_t \end{aligned}$$
(4.46)

with initial condition \(G_0={\bar{v}}\,e^{v{\bar{v}}}\,\). The initial condition is obvious, so let’s check the derivatives. We have

$$\begin{aligned} i\,{\partial \over \partial t}\, G_t= & {} \,-\,\varepsilon \,G_t\;-\;2u\,e^{+i2ut}v{\bar{v}}\, G_t \end{aligned}$$
(4.47)

and one calculates

$$\begin{aligned} \mathcal{L}_0\,G_t= & {} \,-\,\varepsilon \,G_t \nonumber \\ \mathcal{L}_\textrm{int}\,G_t= & {} \,-\,2u\,e^{+i2ut}v{\bar{v}}\, G_t \end{aligned}$$
(5.1)

which verifies (4.38) of the untransformed representation (a).

5 The Two Site Bose-Hubbard Model

In this chapter we consider Bose-Hubbard model with just two lattice sites, 1 and 2. Because of its numerical simplicity, everything can be calculated easily with exact diagonalization, but its highly nontrivial physics, the model exhibits collapse and revivals and has a phase transition between an oscillatory and a self trapping regime, it provides an extremely beautiful test case for any calculation scheme which aims at an efficient and reasonable description of quantum many body systems. In fact, about 80 to 90 percent of the research time has been spent within that model, the generalization to arbitrary dimensions in the end then being more or less straightforward.

The Hamiltonian in the Bargmann-Segal representation is

$$\begin{aligned} h= & {} {\sum }_{i,j=1}^2 \varepsilon _{ij}\,a^+_i a_j\;+\;u\,{\sum }_{j=1}^2 a_j^+a_j^+a_ja_j \nonumber \\= & {} \varepsilon \,\bigl (\,z_1{\partial \over \partial z_2} \,+\, z_2{\partial \over \partial z_1}\,\bigr ) \;+\, u\,\bigl (\; z_1^2\,{\partial ^2\over \partial z_1^2}\,+\, z_2^2\,{\partial ^2\over \partial z_2^2}\;\bigr ) \end{aligned}$$
(5.2)

Thus, the connection to the standard notation with hopping J and interaction strength U is made through

$$\begin{aligned} \varepsilon= & {} -J \end{aligned}$$
(5.3)
$$\begin{aligned} u= & {} U/2 \end{aligned}$$
(5.4)

and we ignore any on-diagonal trapping potentials, that is, we put \(\varepsilon _{jj}=\epsilon _j=0\). Then Theorem 4 of Section 2.2 for coherent states and Theorem 5 of Section 2.3 for number states can be summarized by the following SDE system

$$\begin{aligned} dv_1= & {} \;-\;i\,\varepsilon \, dt\,v_2 \;-\; i\,2u\,dt\, p(v{\bar{v}})\,v_1{\bar{v}}_1 \,v_1\;-\;i\sqrt{2u}\,v_1\, dx_1 \nonumber \\ dv_2= & {} \;-\;i\,\varepsilon \, dt\,v_1 \;-\; i\,2u\,dt\, p(v{\bar{v}})\,v_2{\bar{v}}_2 \,v_2\;-\;i\sqrt{2u}\,v_2\, dx_2 \nonumber \\ d\bar{v}_1= & {} \;+\;i\,\varepsilon \, dt\,\bar{v}_2\;+\;i\, 2u\,dt\, p(v{\bar{v}})\,v_1{\bar{v}}_1 \,{\bar{v}}_1 \;+\;i\sqrt{2u}\,\bar{v}_1\, dy_1 \nonumber \\ d\bar{v}_2= & {} \;+\;i\,\varepsilon \, dt\,\bar{v}_1\;+\;i\, 2u\,dt\, p(v{\bar{v}})\,v_2{\bar{v}}_2 \,{\bar{v}}_2 \;+\;i\sqrt{2u}\,\bar{v}_2\, dy_2 \end{aligned}$$
(5.5)

with \(\,v{\bar{v}}=v_1{\bar{v}}_1+v_2{\bar{v}}_2\,\) and

$$\begin{aligned} p(x)= & {} P'(x)/P(x)\;=\;[\log P]'(x) \end{aligned}$$
(5.6)

with

$$\begin{aligned} P(x)= & {} {\left\{ \begin{array}{ll} e^x &{}\mathrm{\;if\;the\;initial\;state\;is\;a\;coherent\;state} \\ x^{N-1}&{}\mathrm{\;if\;the\;initial\;state\;is\;a\;number\;state} \end{array}\right. } \end{aligned}$$
(5.7)

such that

$$\begin{aligned} p(x)= & {} {\left\{ \begin{array}{ll} 1 &{}\mathrm{\;if\;the\;initial\;state\;is\;a\;coherent\;state} \\ (N-1)/x&{}\mathrm{\;if\;the\;initial\;state\;is\;a\;number\;state} \end{array}\right. } \end{aligned}$$
(5.8)

Recall that the initial coherent and number states were given by

$$\begin{aligned} \psi _0(z_1,z_2)= & {} {\left\{ \begin{array}{ll}\; e^{{\lambda }_1z_1+{\lambda }_2z_2}\,e^{-{|{\lambda }_1|^2+|{\lambda }_2|^2\over 2}} \;=\;e^{{\lambda }z}\,e^{-{|{\lambda }|^2\over 2}} &{}\;\;\mathrm{coherent\;state} \\ \; {1 \over \sqrt{N!N^N}} \;({\lambda }_1z_1+{\lambda }_2z_2)^N\;=\; {({\lambda }z)^N \over \sqrt{N!N^N}} &{}\;\;\mathrm{number\;state} \end{array}\right. } \end{aligned}$$
(5.9)

with

$$\begin{aligned} |{\lambda }|^2\;\;=\;\;|{\lambda }_1|^2+|{\lambda }_2|^2= & {} N \end{aligned}$$
(5.10)

being the total number of particles. The density matrix elements are then given by

$$\begin{aligned} (\psi _t,a_i^+a_j\,\psi _t)_\mathcal{F}= & {} \textsf{E}\bar{\textsf{E}}[\,v_i{\bar{v}}_j\,] \end{aligned}$$
(5.11)

with the \(v_1,v_2\) and \({\bar{v}}_1,{\bar{v}}_2\) given by (5.4) with initial conditions

$$\begin{aligned} v_0\;=\;\begin{pmatrix}v_{1,0}\\ v_{2,0} \end{pmatrix}\;=\;\begin{pmatrix}{\lambda }_1\\ {\lambda }_2 \end{pmatrix},\;\;\;\; {\bar{v}}_0\;=\;\begin{pmatrix}{\bar{v}}_{1,0}\\ {\bar{v}}_{2,0} \end{pmatrix}\;=\;\begin{pmatrix}{\bar{\lambda }}_1\\ {\bar{\lambda }}_2 \end{pmatrix}\end{aligned}$$
(5.12)

Let’s introduce the quadratic quantities

$$\begin{aligned} n_1:= & {} v_1{\bar{v}}_1 \nonumber \\ n_2:= & {} v_2{\bar{v}}_2 \nonumber \\ q:= & {} v_1{\bar{v}}_2 \nonumber \\ {\bar{q}}:= & {} {\bar{v}}_1 v_2 \end{aligned}$$
(5.13)

where the \({\bar{q}}\), as long as no expectation values are taken, is not necessarily the complex conjugate of the q. Then, because of \(\,dv_id{\bar{v}}_i=0\,\), these quantities satisfy the following SDE system. The on-diagonal elements are given by

$$\begin{aligned} dn_1= & {} \;+\;i\,\varepsilon \,dt\,(q-{\bar{q}}) \;-\;i\sqrt{2u}\,n_1\, (dx_1-dy_1) \nonumber \\ dn_2= & {} \;-\;i\,\varepsilon \,dt\,(q-{\bar{q}}) \;-\;i\sqrt{2u}\,n_2\, (dx_2-dy_2) \end{aligned}$$
(5.14)

and for the off-diagonal elements one obtains, with \(n=n_1+n_2=v_1{\bar{v}}_1+v_2{\bar{v}}_2=v{\bar{v}}\),

$$\begin{aligned} dq= & {} \;+\;i\,\varepsilon \, dt\,(n_1-n_2)\;-\; i\,2u\,dt\;p(n)\,(n_1-n_2)\,q \;-\;i\sqrt{2u}\,q\, (dx_1-dy_2) \nonumber \\ d{\bar{q}}= & {} \;-\;i\,\varepsilon \, dt\,(n_1-n_2)\;+\; i\,2u\,dt\;p(n)\,(n_1-n_2)\,{\bar{q}}\;-\;i\sqrt{2u}\,{\bar{q}}\, (dx_2-dy_1) \ \ \ \ \ \ \ \end{aligned}$$
(5.15)

since for example,

$$\begin{aligned} dq= & {} dv_1\,{\bar{v}}_2\;+\;v_1\,d{\bar{v}}_2\;+\;dv_1\,d{\bar{v}}_2 \\= & {} \bigl \{\;-\;i\,\varepsilon \, dt\,v_2 \;-\; i\,2u\,dt\; p(v{\bar{v}})\,v_1{\bar{v}}_1 \,v_1\;-\;i\sqrt{2u}\,v_1\, dx_1\bigr \}\, {\bar{v}}_2\nonumber \\{} & {} \;+\; v_1\,\bigl \{\;i\,\varepsilon \, dt\,\bar{v}_1\;+\;i\, 2u\,dt\; p(v{\bar{v}})\,v_2{\bar{v}}_2 \,{\bar{v}}_2 \;+\;i\sqrt{2u}\,\bar{v}_2\, dy_2 \; \bigr \} \;\;+\;0 \nonumber \\= & {} \;+\;i\,\varepsilon \, dt\,(v_1{\bar{v}}_1-v_2{\bar{v}}_2)\;-\; i\,2u\,dt\;p(v{\bar{v}})\,(v_1{\bar{v}}_1-v_2{\bar{v}}_2)\,v_1{\bar{v}}_2 \;-\;i\sqrt{2u}\,v_1{\bar{v}}_2\, (dx_1-dy_2) \nonumber \end{aligned}$$
(5.16)

Let’s introduce again, now for \(j\in \{1,2\}\),

$$\begin{aligned} d\xi _j:= & {} { dx_j-dy_j \over \sqrt{2}} \nonumber \\ d\eta _j:= & {} { dx_j+dy_j \over \sqrt{2}} \end{aligned}$$
(5.17)

and let’s also put

$$\begin{aligned} d\nu _{12}\;\;:=\;\; \sqrt{2}\,(dx_1-dy_2)= & {} d\xi _1+d\xi _2 +d\eta _1-d\eta _2 \nonumber \\ d\nu _{21}\;\;:=\;\; \sqrt{2}\,(dx_2-dy_1)= & {} d\xi _1+d\xi _2 -d\eta _1+d\eta _2 \end{aligned}$$
(5.18)

Then, with the abbreviations

$$\begin{aligned} n_{12}:= & {} n_1-n_2 \nonumber \\ n:= & {} n_1+n_2 \end{aligned}$$
(5.19)

the system (5.13) and (5.14) looks as follows:

$$\begin{aligned} dn= & {} \;-\;i\sqrt{4u}\,( n_1\, d\xi _1 + n_2\, d\xi _2) \nonumber \\ dn_{12}= & {} \;+\;i\,2\varepsilon \,dt\,(q-{\bar{q}}) \;-\;i\sqrt{4u}\, (n_1\, d\xi _{1}-n_2\, d\xi _{2}) \nonumber \\ dq= & {} \;+\;i\,\varepsilon \, dt\,n_{12}\;-\; i\,2u\,dt\,p(n)\,n_{12}\,q \;-\;i\sqrt{u}\,q\, d\nu _{12} \nonumber \\ d{\bar{q}}= & {} \;-\;i\,\varepsilon \, dt\,n_{12}\;+\; i\,2u\,dt\,p(n)\,n_{12}\,{\bar{q}}\;-\;i\sqrt{u}\,{\bar{q}}\, d\nu _{21} \end{aligned}$$
(5.20)

from which we immediately get \(\,\textsf{E}\bar{\textsf{E}}[\,dn_t\,]=0\,\) or

$$\begin{aligned} \langle n_t\rangle \;\;:=\;\;\textsf{E}\bar{\textsf{E}}[\,n_t\,]= & {} n_0\;\;=\;\;n_{1,0}+n_{2,0}\;\;=\;\;N \;\;\;\; \forall t \end{aligned}$$
(5.21)

Furthermore we have the following exact equations:

$$\begin{aligned} {d\over dt}\langle n_{12,t}\rangle= & {} \;+\;i\,2\varepsilon \,\bigl (\,\langle q_t\rangle \,-\,\langle {\bar{q}}_t\rangle \,\bigr ) \nonumber \\ {d\over dt}\langle q_t\rangle= & {} \;+\;i\,\varepsilon \,\langle n_{12,t}\rangle \;-\; i\,2u\,\bigl \langle \,p(n_t)\,n_{12,t}\,q_t\,\bigr \rangle \nonumber \\ {d\over dt}\langle {\bar{q}}_t\rangle= & {} \;-\;i\,\varepsilon \,\langle n_{12,t}\rangle \;+\; i\,2u\,\bigl \langle \,p(n_t)\,n_{12,t}\,{\bar{q}}_t\,\bigr \rangle \end{aligned}$$
(5.22)

Here we reencounter the typical feature of quantum many body systems, namely, the system is non closed and when trying to close it by deriving SDEs for quantities like \(n_{12,t}\,q_t\) or \(\,p(n_t)\,n_{12,t}\,q_t\,\), we generate higher and higher products which basically corresponds to an expansion of the exponential which generates the collapse and revivals.

5.1 Large N Limit

If we introduce the normalized quantities (so the w’s used in this section are different from the w’s used in chapter 2)

$$\begin{aligned} \rho _1:= & {} n_1/N\;,\qquad \rho _2\;\;:=\;\; n_2/N \nonumber \\ w:= & {} q/N\;, \qquad \ \ {\bar{w}}\;\;:=\;\;{\bar{q}}/N \\{} & {} \ \ g\;\;:=\;\;uN \nonumber \end{aligned}$$
(5.23)

and

$$\begin{aligned} \rho _{12}:= & {} \rho _1-\rho _2 \nonumber \\ \rho:= & {} \rho _1+\rho _2 \end{aligned}$$
(5.24)

the system (5.19) is equivalent to (with \(\,p(n)=p(N\rho )\,\), we write p(n) for brevity)

$$\begin{aligned} d\rho= & {} \;-\;i\,\sqrt{4g/N}\,( \rho _1\, d\xi _1 + \rho _2\, d\xi _2) \nonumber \\ d\rho _{12}= & {} \;+\;i\,2\varepsilon \,dt\,(w-{\bar{w}}) \;-\;i\,\sqrt{4g/N}\, (\rho _1\, d\xi _{1}-\rho _2\, d\xi _{2}) \nonumber \\ dw= & {} \;+\;i\,\varepsilon \, dt\,\rho _{12}\;-\; i\,2g\,dt\,p(n)\,\rho _{12}\,w \;-\;i\,\sqrt{g/N}\,w\, d\nu _{12} \nonumber \\ d{\bar{w}}= & {} \;-\;i\,\varepsilon \, dt\,\rho _{12}\;+\;i\,2g\,dt\,p(n)\,\rho _{12}\,{\bar{w}}\;-\;i\,\sqrt{g/N}\,{\bar{w}}\, d\nu _{21} \end{aligned}$$
(5.25)

In the limit \(\,N\rightarrow \infty \,\) with \(\,g=uN\,\) held fixed, the diffusive part vanishes and we obtain the ODE system

$$\begin{aligned} \dot{\rho }= & {} 0 \nonumber \\ \dot{\rho }_{12}= & {} \;+\;i\,2\varepsilon \,(w-{\bar{w}}) \nonumber \\ \dot{w}= & {} \;+\;i\,\varepsilon \,\rho _{12}\;-\; i\,2g\,\rho _{12}\,w \nonumber \\ \dot{{\bar{w}}}= & {} \;-\;i\,\varepsilon \,\rho _{12}\;+\; i\,2g\,\rho _{12}\,{\bar{w}}\end{aligned}$$
(5.26)

Here we can ignore the \(\,p(n_t)=p(N\rho _t)\,\) also for number states since we have \(\,p(N\rho _t)=(N-1)/(N\rho _t)=(N-1)/N\rightarrow 1\,\) because of \(\,\rho _t=1\,\) for all t. Now let’s put all N particles on lattice site 1, that is, we choose the initial conditions

$$\begin{aligned} (\rho _1,\rho _2,w,{\bar{w}})_0=(1,0,0,0) \end{aligned}$$
(5.27)

The quantity \({\bar{w}}_t\) is now the true complex conjugate of \(w_t\). We are left with the two equations

$$\begin{aligned} \dot{\rho }_{12}= & {} \;-\;4\varepsilon \;\textrm{Im}\,w \end{aligned}$$
(5.28)
$$\begin{aligned} \dot{w}= & {} \;+\;i\,\varepsilon \,\rho _{12}\;-\; i\,2g\,\rho _{12}\,w \end{aligned}$$
(5.29)

Equation (5.28) is solved by

$$\begin{aligned} w_t= & {} {\varepsilon \over 2g}\;+\; \bigl ( w_0 \,-\,{\varepsilon \over 2g}\bigr )\; e^{\,-2ig\int _0^t \rho _{12,s}ds} \nonumber \\&\buildrel w_0=0\over =&{\varepsilon \over 2g}\;-\; {\varepsilon \over 2g}\; e^{\,-2ig\int _0^t \rho _{12,s}ds} \end{aligned}$$
(5.30)

We obtain

$$\begin{aligned} \textrm{Im}\,w_t= & {} \,+\,{\varepsilon \over 2g}\,\sin \bigl [\, 2g\int _0^t \rho _{12,s} ds\,\bigr ] \end{aligned}$$
(5.31)

Putting this into (5.27) gives

$$\begin{aligned} \dot{\rho }_{12,t}= & {} \;-\;4\varepsilon \,\textrm{Im}\,w_t\;\;=\;\; \,-\,{2\varepsilon ^2 \over g}\,\sin \bigl [\, 2g\int _0^t \rho _{12,s} ds\,\bigr ] \end{aligned}$$
(5.32)

Thus, the quantity

$$\begin{aligned} \varphi _t:= & {} 2g\int _0^t \rho _{12,s} ds \end{aligned}$$
(5.33)

is the solution of the second order equation

$$\begin{aligned} \ddot{\varphi }_t \;+\; 4\varepsilon ^2 \sin \varphi _t= & {} 0 \end{aligned}$$
(5.34)

which is just the equation of motion for the mathematical pendulum. We have the following initial conditions:

$$\begin{aligned} \varphi _0= & {} 0 \nonumber \\ \dot{\varphi }_0= & {} 2g\,\rho _{12,0}\;=\;2g\,(1-0)\;\;=\;\;2g \end{aligned}$$
(5.35)

with total energy

$$\begin{aligned} E= & {} {\dot{\varphi }_t^2\over 2}\,-\,4\varepsilon ^2 \cos \varphi _t\;\;=\;\;{\dot{\varphi }_0^2\over 2}\,-\,4\varepsilon ^2 \cos \varphi _0 \;\;=\;\;2g^2\,-\,4\varepsilon ^2 \end{aligned}$$
(5.36)

The potential energy at \(\,\varphi =\pi \,\) is \(\,E_\textrm{pot}=+4\varepsilon ^2\,\). We have rollovers if the total energy is bigger than that, that is, if \(\,2g^2\,-\,4\varepsilon ^2 > +4\varepsilon ^2\,\) or

$$\begin{aligned} g^2> & {} (2\varepsilon )^2 \;. \end{aligned}$$
(5.37)

Numerical Test

Let’s make a numerical check. We choose the following values: \(\,\varepsilon =1\,\) and

$$\begin{aligned} N\in & {} \{\,2500\,,\,5000\,,\,10000\,,\,20000\,\} \end{aligned}$$
(5.38)
$$\begin{aligned} g\in & {} \{\,0.5\,,\,1.0\,,\,1.8\,,\,2.2\,,\,3.0\,,\,6.0\,\} \end{aligned}$$
(5.39)

and calculate the quantity

$$\begin{aligned} \rho _{1,t}\;\;=\;\;n_{1,t}/N= & {} (\psi _t,a_1^+a_1\,\psi _t)_\mathcal{F}\;/\;N \end{aligned}$$
(5.40)

in two different ways: First, by exact diagonalization. There we have to use the different values for N given by (5.37). Second, by simulating the ODE system 5.33 for the mathematical pendulum and calculating \(\,\rho _{1,t}\,\) through

$$\begin{aligned} \rho _{1,t}= & {} {1\over 2}\bigl (\, 1\,+\, {\dot{\varphi }_t \over 2g}\,\bigr ) \end{aligned}$$
(5.41)

This is the large N limit and accordingly no N enters the calculation, but only a value for g. We obtain the following results as displayed in Figs. 1 and 2 below. The red line is the ODE solution and the dots come from exact diagonalization.

Fig. 1
figure 1

Exact diagonalization results vs. ODE solution for the quantity \(\langle n_{1,t}\rangle /N\)

Fig. 2
figure 2

Exact diagonalization results vs. ODE solution for the quantity \(\langle n_{1,t}\rangle /N\)

The closer the value of g approaches 2, the larger N has to be chosen in order to numerically reach the true \(N=+\infty \) limit. The different colors of the dots represent the different values of N, with the obvious ordering of orange, green, light blue and dark blue for increasing values of N. The fact that \(\,g=2\varepsilon =2\,\) marks the transition point is also visualized through the following picture,

figure b

which shows the quantity (5.40) for \(g=1.99\) in black and for \(g=2.01\) in red. The fact that the mathematical pendulum shows up in the dynamics of the two site Bose-Hubbard model has been observed by several authors, for example in refs [9,10,11,12]. The very beautiful thesis of Lena Simon [13] also provides a detailed discussion of the dynamics.

Equivalence to Quartic Double Well Potential

Actually we can obtain \(n_{12,t}\) or \(\rho _{12,t}\) also as a solution of a classical particle moving in a quartic double well potential. Recall the ODE system (5.27),(5.28),

$$\begin{aligned} \dot{\rho }_{12}= & {} \;-\;4\varepsilon \;\textrm{Im}\,w \nonumber \\ \dot{w}= & {} \;+\;i\,\varepsilon \,\rho _{12}\;-\; i\,2g\,\rho _{12}\,w \end{aligned}$$
(5.42)

We have

$$\begin{aligned} \ddot{\rho }_{12}= & {} \;-\;4\varepsilon \;\textrm{Im}\,\dot{w} \nonumber \\= & {} \; -\,4\varepsilon ^2\rho _{12}\;+\; 8g\varepsilon \,\rho _{12}\,\textrm{Re}\,w \end{aligned}$$
(5.43)

and

$$\begin{aligned} \textrm{Re}\,\dot{w}= & {} \;+\; 2g\,\rho _{12}\,\textrm{Im}\,w \end{aligned}$$
(5.44)

or

$$\begin{aligned} 4\varepsilon \,\textrm{Re}\,\dot{w}= & {} \;-\; 2g\,\rho _{12}\,\dot{\rho }_{12} \end{aligned}$$
(5.45)

which gives, with initial conditions \(w_0=0\) and \(\rho _{12,0}=1\),

$$\begin{aligned} 4\varepsilon \,\textrm{Re}\, w_t= & {} g(\rho _{12,0}^2-\rho _{12,t}^2)\;\;=\;\;g(1-\rho _{12,t}^2) \end{aligned}$$
(5.46)

Thus,

$$\begin{aligned} \ddot{\rho }_{12,t}\; +\,4\varepsilon ^2\rho _{12,t}= & {} \;+\; 2g\,\rho _{12,t}\,4\varepsilon \,\textrm{Re}\,w_t\, \nonumber \\= & {} \;+\; 2g^2\,\rho _{12,t}\,(1-\rho _{12,t}^2) \end{aligned}$$
(5.47)

or

$$\begin{aligned} \ddot{\rho }_{12,t}\; +\;(4\varepsilon ^2-2g^2)\rho _{12,t}\;+\;2g^2\,\rho _{12,t}^3= & {} 0 \end{aligned}$$
(5.48)

with initial conditions \(\rho _{12,0}=1\) and \(\dot{\rho }_{12,0}=0\,\). Let’s summarize in the following

Theorem 7

The mathematical pendulum

$$\begin{aligned} \ddot{\varphi }_t\;+\;4\varepsilon ^2\,\sin \varphi _t= & {} 0 \end{aligned}$$
(5.49)

with \(\varphi _0=0\) and \(\dot{\varphi }_0=2g\) is equivalent to the cubic equation

$$\begin{aligned} \ddot{x}_{t}\; +\;(4\varepsilon ^2-2g^2)\,x_{t}\;+\;2g^2\,x_{t}^3= & {} 0 \end{aligned}$$
(5.50)

with \(x_{0}=1\) and \(\dot{x}_{0}=0\) through the transformation

$$\begin{aligned} \varphi _t\;\;=\;\; 2g\int _0^t x_{s}\;\text {ds} \;\;\;\;\;\; \Leftrightarrow \;\;\;\;\;\; x_{t}\;\;=\;\;{1 \over 2g}\; \dot{\varphi }_t \end{aligned}$$
(5.51)

Proof

Since this is a stand-alone result independent of the rest of the paper, let us also write down a stand-alone proof independent of the earlier derivations. We have

$$\begin{aligned}{ 2g\,\bigl [\;\ddot{x}_{t}\; +\;(4\varepsilon ^2-2g^2)\,x_{t}\;+\;2g^2\,x_{t}^3\;\bigr ]\;\;=\;\; {d\over dt}\,\ddot{\varphi }_t \;+\;(4\varepsilon ^2-2g^2)\,\dot{\varphi }_t \;+\; {\dot{\varphi }_t^3\over 2}\; } \\= & {} {d\over dt}\bigl [\,\ddot{\varphi }_t\,+\,4\varepsilon ^2\sin \varphi _t\,\bigr ]\;+\;\dot{\varphi }_t\,\bigl [\, {\dot{\varphi }_t^2\over 2}\;-\; 4\varepsilon ^2\cos \varphi _t \;-\;(2g^2-4\varepsilon ^2)\,\bigr ] \\= & {} {d\over dt}\bigl [\,\ddot{\varphi }_t\,+\,4\varepsilon ^2\sin \varphi _t\,\bigr ]\;+\;\dot{\varphi }_t\,\int _0^t\,{d\over ds}\bigl [\, {\dot{\varphi }_s^2\over 2}\,-\, 4\varepsilon ^2\cos \varphi _s\,\bigr ]\,ds \\= & {} {d\over dt}\bigl [\,\ddot{\varphi }_t\,+\,4\varepsilon ^2\sin \varphi _t\,\bigr ]\;+\;\dot{\varphi }_t\,\int _0^t\,\bigl [\, \ddot{\varphi }_s\,+\, 4\varepsilon ^2\sin \varphi _s\,\bigr ]\,\dot{\varphi }_s\,ds \end{aligned}$$

Thus, if (5.48) holds, then also (5.49) is fulfilled. On the other hand, if (5.49) holds, then the quantity \(\,F_t\,:=\,\ddot{\varphi }_t\,+\,4\varepsilon ^2\sin \varphi _t\,\) satisfies the equation

$$\begin{aligned}\dot{F}_t\;+\;4g^2 x_t \int _0^t F_s\,x_s\, ds= & {} 0 \end{aligned}$$

or

$$\begin{aligned}x_t\,\ddot{F}_t\;-\;\dot{x}_t\,\dot{F}_t \;+\;4g^2 x_t^3\, F_t= & {} 0 \end{aligned}$$

which is a linear homogeneous second order equation. Since we have the initial conditions \(F_0=0\) and, using (5.49) again, also \(\dot{F}_0=0\), \(F_t\) has to vanish identically. \(\square \)

The transition between the oscillatory and the self trapping regime is intuitive for the mathematical pendulum, let’s try to understand this also by using the cubic equation. The total energy for (5.49) is

$$\begin{aligned} E= & {} {\dot{\rho }_{12,t}^2\over 2} \;+\;(2\varepsilon ^2-g^2)\rho _{12,t}^2\;+\;{g^2\over 2}\,\rho _{12,t}^4\;\;=\;\; {\,4\varepsilon ^2 \,-\, g^2\,\over 2} \end{aligned}$$
(5.52)

The potential energy is

$$\begin{aligned} E_\textrm{pot}(\rho _{12})= & {} (2\varepsilon ^2-g^2)\rho _{12,t}^2\;+\;{g^2\over 2}\,\rho _{12,t}^4 \end{aligned}$$
(5.53)

and has stationary points at \(\rho _{12}=0\) and

$$\begin{aligned} 4\varepsilon ^2\,-\,2g^2\;+\;2\,g^2\,\rho _{12,t}^2= & {} 0 \end{aligned}$$
(5.54)

or

$$\begin{aligned} \rho _{12,t}^2= & {} 1\;-\;{2\varepsilon ^2\over g^2}\;\;<\;\; 1 \end{aligned}$$
(5.55)

Thus, for \(g^2>2\varepsilon ^2\) we have a double well potential which, for \(E<0\) or \(g^2>4\varepsilon ^2\), is sufficiently deep such that \(\rho _{12,t}\) cannot escape the right well when starting at \(\rho _{12,0}=1\) with \(\dot{\rho }_{12,0}=0\,\). Since, \(\rho _{12,t}\) starts at 1 with a negative energy and the potential energy at \(\rho _{12}=0\) is zero such that the energy required to cross this point is \(\,{\dot{\rho }_{12,t}^2/ 2} >0\,\) which is not available.

Obviously, the mathematical pendulum has no collapse and revivals, so in order to see these, we have to take the diffusive part of the SDE system into account. Before we do this, let’s have a look at PDE representations.

5.2 PDE Representations

In chapter 3, we used the \(v_j,{\bar{v}}_j\) as the basic variables to obtain PDE representations for the density matrix elements. Now, it is very instructive to see the corresponding PDE representations if the quadratic quantities

$$\begin{aligned} (n_1,n_2,q,{\bar{q}})= & {} (v_1{\bar{v}}_1,v_2{\bar{v}}_2,v_1{\bar{v}}_2,v_2{\bar{v}}_1) \end{aligned}$$
(5.56)

are used directly as variables. To this end recall the SDE system (5.13) and (5.14) for the quadratic quantities from the last section,

$$\begin{aligned} dn_1= & {} \;+\;i\,\varepsilon \,(q-{\bar{q}})\,dt \;-\;i\sqrt{4u}\,n_1\, d\xi _1 \nonumber \\ dn_2= & {} \;-\;i\,\varepsilon \,(q-{\bar{q}})\,dt \;-\;i\sqrt{4u}\,n_2\, d\xi _2 \nonumber \\ dq= & {} \;+\;i\,\varepsilon \,n_{12}\,dt\;-\; i\,2u\, p(n)\,n_{12}\,q\,dt \;-\;i\sqrt{u}\,q\, d\nu _{12} \nonumber \\ d{\bar{q}}= & {} \;-\;i\,\varepsilon \,n_{12}\,dt\;+\; i\,2u\, p(n)\,n_{12}\,{\bar{q}}\,dt \;-\;i\sqrt{u}\,{\bar{q}}\, d\nu _{21} \end{aligned}$$
(5.57)

with \(n=n_1+n_2=v{\bar{v}}\), \(\,n_{12}=n_1-n_2\,\) and

$$\begin{aligned} p(x)= & {} P'(x)/P(x)\;=\;[\log P]'(x) \end{aligned}$$
(5.58)

with

$$\begin{aligned} P(x)= & {} {\left\{ \begin{array}{ll} e^x &{}\mathrm{\;if\;the\;initial\;state\;is\;a\;coherent\;state} \\ x^{N-1}&{}\mathrm{\;if\;the\;initial\;state\;is\;a\;number\;state} \end{array}\right. } \end{aligned}$$
(5.59)

such that

$$\begin{aligned} p(x)= & {} {\left\{ \begin{array}{ll} 1 &{}\mathrm{\;if\;the\;initial\;state\;is\;a\;coherent\;state} \\ (N-1)/x&{}\mathrm{\;if\;the\;initial\;state\;is\;a\;number\;state} \end{array}\right. } \end{aligned}$$
(5.60)

Recall the abbreviations

$$\begin{aligned} d\xi _1= & {} {(dx_1-dy_1)/ \sqrt{2}} \nonumber \\ d\xi _2= & {} {(dx_2-dy_2)/ \sqrt{2}} \nonumber \\ d\nu _{12}= & {} \sqrt{2}\;(dx_1-dy_2) \nonumber \\ d\nu _{21}= & {} \sqrt{2}\;(dx_2-dy_1) \end{aligned}$$
(5.61)

To write down a PDE representation which is again obtained as a Kolmogorov backward equation, recall the logic of Appendices A.3 and A.4, we need to determine the differential operator A which is associated to the SDE system (5.56). To do this, we need the following identities which determine the second order part of A:

$$\begin{aligned} (d\xi _1)^2\;=\;(d\xi _2)^2\;=\;(d\nu _{12})^2\;=\;(d\nu _{21})^2\;=\;d\xi _1d\xi _2\;=\;d\nu _{12}d\nu _{21}\;=\;0 \end{aligned}$$
(5.62)

and

$$\begin{aligned} d\xi _1d\nu _{12}= & {} (dx_1)^2\;=\;+\,i\,dt \nonumber \\ d\xi _1d\nu _{21}= & {} (dy_1)^2\;=\;-\,i\,dt \nonumber \\ d\xi _2d\nu _{12}= & {} (dy_2)^2\;=\;-\,i\,dt \nonumber \\ d\xi _2d\nu _{21}= & {} (dx_2)^2\;=\;+\,i\,dt \end{aligned}$$
(5.63)

Now let

$$\begin{aligned}F= & {} F(n_1,n_2,q,{\bar{q}}) \end{aligned}$$

by any function. We plug in the \(n_{1t},n_{2,t}\) and \(q_t,{\bar{q}}_t\) from the SDE system above and calculate the dF with the Ito lemma:

$$\begin{aligned} dF= & {} {\partial F\over \partial n_1}\,dn_1\;+\;{\partial F\over \partial n_2}\,dn_2\;+\;{\partial F\over \partial q}\,dq\;+\;{\partial F\over \partial {\bar{q}}}\,d{\bar{q}}\\{} & {} \;+\; {\partial ^2 F\over \partial n_1\partial q}\,dn_1dq \;+\;{\partial ^2 F\over \partial n_1\partial {\bar{q}}}\,dn_1d{\bar{q}}\;+\;{\partial ^2 F\over \partial n_2\partial q}\,dn_2dq \;+\;{\partial ^2 F\over \partial n_2\partial {\bar{q}}}\,dn_2d{\bar{q}}\nonumber \\= & {} \;+\;i\varepsilon dt\,(q-{\bar{q}})\, {\partial F\over \partial n_1}\;-\;i\varepsilon dt\,(q-{\bar{q}}) \,{\partial F\over \partial n_2} \nonumber \\{} & {} \;+\;\bigl \{\,i dt\,\varepsilon \,n_{12,t} \,-\,idt\,2u\,p(n)\,n_{12,t}\,q_t \,\bigr \}\, {\partial F\over \partial q} \;-\;\bigl \{\,i dt\,\varepsilon \,n_{12,t} \;-\;idt\,2u\,p(n)\,n_{12,t}\,{\bar{q}}_t \,\bigr \}\,{\partial F\over \partial {\bar{q}}} \nonumber \\{} & {} \;-\;2u\,\Bigl \{\; {\partial ^2 F\over \partial n_1\partial q}\,n_1q \,d\xi _1d\nu _{12} \;+\;{\partial ^2 F\over \partial n_1\partial {\bar{q}}}\,n_1{\bar{q}}\,d\xi _1d\nu _{21} \nonumber \\{} & {} \;+\;{\partial ^2 F\over \partial n_2\partial q}\,n_2q\, d\xi _2d\nu _{12} \;+\;{\partial ^2 F\over \partial n_2\partial {\bar{q}}}\,n_2{\bar{q}}\, d\xi _2d\nu _{21} \;\Bigr \} \;\;+\;\; \textrm{diffusive} \nonumber \end{aligned}$$
(5.64)

or, using (5.62),

$$\begin{aligned} dF= & {} \;+\;i\varepsilon \,(q-{\bar{q}})\, \bigl ( {\partial F\over \partial n_1}\,- \,{\partial F\over \partial n_2} \bigr )\,dt \;+\;i \,\varepsilon \,(n_1-n_2) \, \bigl ( {\partial F\over \partial q} \,- \,{\partial F\over \partial {\bar{q}}} \bigr ) \,dt \\{} & {} \;-\;i\,2u\,p(n)\,(n_1-n_2)\,\bigl ( q \, {\partial F\over \partial q} \,-\,{\bar{q}}\,{\partial F\over \partial {\bar{q}}} \bigr )\,dt \nonumber \\{} & {} \;-\;i\,2u\,\bigl (\, {\partial ^2 F\over \partial n_1\partial q}\,n_1q \,-\,{\partial ^2 F\over \partial n_1\partial {\bar{q}}}\,n_1{\bar{q}}\,-\,{\partial ^2 F\over \partial n_2\partial q}\,n_2q \,+\,{\partial ^2 F\over \partial n_2\partial {\bar{q}}}\,n_2{\bar{q}}\,\bigr )\,dt \;\;+\;\; \textrm{diffusive} \nonumber \\=: & {} AF\,dt \;\;+\;\; \textrm{diffusive} \end{aligned}$$
(5.65)

Thus, any expectation

$$\begin{aligned} F:= & {} \textsf{E}\bar{\textsf{E}}_0^t[\,f(n_{1,t},n_{2,t},q_t,{\bar{q}}_t)\,\bigr ]\;\;=\;\;F_t(n_{1,0},n_{2,0},q_0,{\bar{q}}_0) \end{aligned}$$
(5.66)

considered as a function of its initial values, has to be a solution of (again, we drop the zero subscripts on the right hand side of (5.66))

$$\begin{aligned} {\partial F \over \partial t}\;=\;AF= & {} \;+\;i\varepsilon \,(q-{\bar{q}})\, \bigl ( {\partial F\over \partial n_1}\,- \,{\partial F\over \partial n_2} \bigr ) \;+\;i \,\varepsilon \,(n_1-n_2) \, \bigl ( {\partial F\over \partial q} \,- \,{\partial F\over \partial {\bar{q}}} \bigr ) \nonumber \\{} & {} \;-\;i\,2u\,p(n)\,(n_1-n_2)\,\bigl ( q \, {\partial F\over \partial q} \,-\,{\bar{q}}\,{\partial F\over \partial {\bar{q}}} \bigr ) \nonumber \\{} & {} \;-\;i\,2u\,\bigl (\, {\partial ^2 F\over \partial n_1\partial q}\,n_1q \,-\,{\partial ^2 F\over \partial n_1\partial {\bar{q}}}\,n_1{\bar{q}}\,-\,{\partial ^2 F\over \partial n_2\partial q}\,n_2q \,+\,{\partial ^2 F\over \partial n_2\partial {\bar{q}}}\,n_2{\bar{q}}\,\bigr ) \nonumber \\=: & {} \;-\;i\,(\mathcal{L}_\varepsilon \,+\,\mathcal{L}_u\,)F \end{aligned}$$
(5.67)

with initial condition \(\,F_0=f\,\) and differential operators

$$\begin{aligned} \mathcal{L}_\varepsilon:= & {} \;-\; \varepsilon \,\Bigl \{\; (q-{\bar{q}})\, \bigl ({\partial \over \partial n_1}-{\partial \over \partial n_2}\bigr ) \;+\; (n_1-n_2) \,\bigl ( {\partial \over \partial q} - {\partial \over \partial {\bar{q}}}\bigr )\;\Bigr \} \end{aligned}$$
(5.68)
$$\begin{aligned} \mathcal{L}_u:= & {} \;\!+\!\;2u \,\Bigl \{\; \bigl (\, n_1 \, {\partial \over \partial n_1} \,\!-\!\,n_2 \,{\partial \over \partial n_2} \,\bigr ) \bigl (\, q \, {\partial \over \partial q} \,\!-\!\,{\bar{q}}\,{\partial \over \partial {\bar{q}}} \,\bigr ) \;+\; p(n)\,(n_1-n_2)\bigl (\, q \, {\partial \over \partial q} \,\!-\!\,{\bar{q}}\,{\partial \over \partial {\bar{q}}} \,\bigr ) \;\Bigr \} \ \ \ \ \ \ \end{aligned}$$
(5.69)

This proves part (a) of the following

Theorem 8

Consider the two site Bose-Hubbard model with Hamiltonian

$$\begin{aligned} h= & {} \varepsilon \,(\,a^+_1 a_2\,+\,a_2^+a_1)\;+\;u\,(\, a_1^+a_1^+a_1a_1\,+\, a_2^+a_2^+a_2a_2\, ) \nonumber \\= & {} \varepsilon \,\bigl (\,z_1{\partial \over \partial z_2} \,+\, z_2{\partial \over \partial z_1}\,\bigr ) \;+\, u\,\bigl (\, z_1^2\,{\partial ^2\over \partial z_1^2}\,+\, z_2^2\,{\partial ^2\over \partial z_2^2}\;\bigr ) \end{aligned}$$
(5.70)

and initial state

$$\begin{aligned} \psi _0(z_1,z_2)= & {} {\left\{ \begin{array}{ll} e^{\,{\lambda }_1z_1\,+\,{\lambda }_2z_2}\; e^{\,-\,{|{\lambda }_1|^2\,+\,|{\lambda }_2|^2\over 2}} &{}\mathrm{\;if\;coherent\;is\;chosen} \\ {1 \over \sqrt{N!N^N}} \;({\lambda }_1z_1 +{\lambda }_2z_2 )^N &{}\mathrm{\;if\;number\;is\;chosen} \end{array}\right. } \end{aligned}$$
(5.71)

with a total number of \(N=|{\lambda }_1|^2+|{\lambda }_2|^2\) particles. Let P be the function of one variable given by

$$\begin{aligned} P(x)= & {} {\left\{ \begin{array}{ll} e^x &{}\mathrm{\;if\;coherent\;is\;chosen} \\ x^{N-1}&{}\mathrm{\;if\;number\;is\;chosen} \end{array}\right. } \end{aligned}$$
(5.72)

and let \(p(x)=P'(x)/P(x)\). Then the following statements hold:

  • a) The expected number of particles \(\,\langle n_{1,t}\rangle =\textsf{E}\bar{\textsf{E}}[n_{1,t}]=(\psi _t,a_1^+a_1\psi _t)_{\mathcal{F}}\,\) at lattice site 1 can be written as

    $$\begin{aligned} \langle n_{1,t}\rangle= & {} e^{-it(\mathcal{L}_\varepsilon \,+\,\mathcal{L}_u\,)}\,n_1\;{\bigr |_{\,(n_1,n_2,q,{\bar{q}})\,=\,(\,|{\lambda }_1|^2,\,|{\lambda }_2|^2,\,{\lambda }_1{\bar{\lambda }}_2\,,\,{\bar{\lambda }}_1{\lambda }_2\,)} } \end{aligned}$$
    (5.73)

    with the differential operators \(\mathcal{L}_\varepsilon \) and \(\mathcal{L}_u\) given by (5.68) and (5.69) above.

  • b) The actions of \(e^{-it\mathcal{L}_\varepsilon }\) and \(e^{-it\mathcal{L}_u}\) are as follows: For \(e^{-it\mathcal{L}_\varepsilon }\) we obtain

    $$\begin{aligned} (e^{-it\mathcal{L}_\varepsilon }F)(n_1,n_2,q,{\bar{q}})= & {} F\bigl (\, R_t\,(n_1,n_2,q,{\bar{q}})^T \,\bigr ) \end{aligned}$$
    (5.74)

    where \(F=F(n_1,n_2,q,{\bar{q}})\) is an arbitrary function and \(R_t\) is the \(4\times 4\) matrix

    $$\begin{aligned} R_t= & {} \begin{pmatrix}\cos ^2\varepsilon t &{} \sin ^2\varepsilon t &{} +i\sin \varepsilon t\cos \varepsilon t &{} -i\sin \varepsilon t\cos \varepsilon t \\ \sin ^2\varepsilon t &{}\cos ^2\varepsilon t &{} -i\sin \varepsilon t\cos \varepsilon t &{} +i\sin \varepsilon t\cos \varepsilon t \\ +i\sin \varepsilon t\cos \varepsilon t &{} -i\sin \varepsilon t\cos \varepsilon t &{}\cos ^2\varepsilon t &{} \sin ^2\varepsilon t \\ -i\sin \varepsilon t\cos \varepsilon t &{} +i\sin \varepsilon t\cos \varepsilon t &{} \sin ^2\varepsilon t &{}\cos ^2\varepsilon t \end{pmatrix}\end{aligned}$$
    (5.75)

    For \(e^{-it\mathcal{L}_u}\) we find:

    $$\begin{aligned} { e^{-it\mathcal{L}_u}\,\bigl \{\, G(n_1,n_2)\,q^b\,{\bar{q}}^{\,{\bar{b}}}\,\bigr \} \;\;=\;\; } \\{} & {} G\bigl (\,e^{-i2u t\,(b-{\bar{b}})}n_1 \,, \, e^{+i2u t\,(b-{\bar{b}})}n_2\,\bigr ) \;\times \; { P(\,e^{-i2u t\,(b-{\bar{b}})}n_1 \, + \, e^{+i2u t\,(b-{\bar{b}})}n_2\,) \over P(n_1+n_2)}\, \times \,q^b\,{\bar{q}}^{\,{\bar{b}}} \nonumber \end{aligned}$$
    (5.76)

    where \(G=G(n_1,n_2)\) is an arbitrary function and \(b,{\bar{b}}\) are arbitrary natural numbers.

Proof

It remains to prove part (b). If we make the Ansatz

$$\begin{aligned} (e^{-it\mathcal{L}_\varepsilon }F)(n_1,n_2,q,{\bar{q}})= & {} F\bigl (\, e^{-itA}\, (n_1,n_2,q,{\bar{q}})^T \,\bigr ) \end{aligned}$$
(5.77)

and take the time derivative, we find

$$\begin{aligned} A= & {} -\varepsilon \begin{pmatrix}0 &{} \sigma \\ \sigma &{} 0 \end{pmatrix}\;\; \in \;\;{\mathbb R}^{4\times 4} \end{aligned}$$
(5.78)

with

$$\begin{aligned} \sigma:= & {} \begin{pmatrix}+1 &{} -1 \\ -1 &{} + 1 \end{pmatrix}\;\;\in \;\; {\mathbb R}^{2\times 2} \end{aligned}$$
(5.79)

Since

$$\begin{aligned} A^{2k}\;\;=\;\;(-\varepsilon )^{2k}\,\begin{pmatrix}\sigma ^{2k} &{} 0 \\ 0 &{} \sigma ^{2k} \end{pmatrix},\;\;\;\;\;\; A^{2k+1}\;\;=\;\;(-\varepsilon )^{2k+1}\,\begin{pmatrix}0&{} \sigma ^{2k+1} \\ \sigma ^{2k+1} &{} 0 \end{pmatrix}\end{aligned}$$
(5.80)

and because of \(\,\sigma ^n=2^{n-1}\sigma =2^n\,{\sigma \over 2}\,\) for \(n\ge 1\), we obtain

$$\begin{aligned} e^{-itA}= & {} Id \;\;+\;\;\sum \limits _{k=1}^\infty { (+2it\varepsilon )^{2k}\over (2k)!} \;{1\over 2}\begin{pmatrix}\sigma &{} 0 \\ 0 &{} \sigma \end{pmatrix}\;\;+\;\; \sum \limits _{k=0}^\infty { (+2it\varepsilon )^{2k+1}\over (2k+1)!} \;{1\over 2}\begin{pmatrix}0&{} \sigma \\ \sigma &{} 0 \end{pmatrix}\nonumber \\= & {} \begin{pmatrix}Id &{} 0 \\ 0 &{} Id \end{pmatrix}\;\;+\;\;{\cos (2\varepsilon t)-1\over 2}\begin{pmatrix}\sigma &{} 0 \\ 0 &{} \sigma \end{pmatrix}\;\;+\;\;i\;{\sin (2\varepsilon t)\over 2}\begin{pmatrix}0&{} \sigma \\ \sigma &{} 0 \end{pmatrix}\end{aligned}$$
(5.81)

which coincides with \(\,R_t\,\) since \(\,\cos ^2\varepsilon t=(1+\cos 2\varepsilon t)/2\,\) and \(\,\sin ^2\varepsilon t=(1-\cos 2\varepsilon t)/2\,\).

The action of \(e^{-it\mathcal{L}_u}\) we calculate by evaluating the Fresnel expectation directly. That is, we write down the \(\,\varepsilon =0\,\) SDE system

$$\begin{aligned} dn_1= & {} \;-\;i\sqrt{4u}\,n_1\, d\xi _1 \nonumber \\ dn_2= & {} \;-\;i\sqrt{4u}\,n_2\, d\xi _2 \nonumber \\ dq= & {} \;-\; i\,2u\,dt\, p(n)\,n_{12}\,q \;-\;i\sqrt{u}\,q\, d\nu _{12} \nonumber \\ d{\bar{q}}= & {} \;+\; i\,2u\,dt\, p(n)\,n_{12}\,{\bar{q}}\;-\;i\sqrt{u}\,{\bar{q}}\, d\nu _{21} \end{aligned}$$
(5.82)

which is solved by

$$\begin{aligned} n_{1,t}= & {} n_1\,e^{\,-\,i\sqrt{4u}\int _0^t d\xi _{1,s} } \nonumber \\ n_{2,t}= & {} n_2\,e^{\,-\,i\sqrt{4u}\int _0^t d\xi _{2,s} } \end{aligned}$$
(5.83)

and

$$\begin{aligned} q_{t}= & {} q\,e^{\,-\,i2u\int _0^t p(n_s)\,n_{12,s}\,ds\,-\,i\sqrt{u}\int _0^t d\nu _{12,s} } \nonumber \\ {\bar{q}}_{t}= & {} {\bar{q}}\,e^{\,+\,i2u\int _0^t p(n_s)\,n_{12,s}\,ds\,-\,i\sqrt{u}\int _0^t d\nu _{21,s} } \end{aligned}$$
(5.84)

We have to evaluate

$$\begin{aligned} { e^{-it\mathcal{L}_u}\,\bigl \{\, G(n_1,n_2)\,q^b\,{\bar{q}}^{\,{\bar{b}}}\,\bigr \}\;\;=\;\;\textsf{E}\bar{\textsf{E}}\bigl [\; G(n_{1,t},n_{2,t})\,q_t^b\,{\bar{q}}_t^{\,{\bar{b}}}\;\bigr ] } \nonumber \\= & {} \textsf{E}\bar{\textsf{E}}\Bigl [\, G(n_{1,t},n_{2,t})\,e^{\,-\,i2u(b-{\bar{b}})\int _0^t p(n_s)\,n_{12,s}\,ds } \,e^{ \,-\,i\sqrt{u}\,\int _0^t[\,b\,d\nu _{12,s}\,+ \,{\bar{b}}\,d\nu _{21,s}\,] } \,\Bigr ] \,q^b\,{\bar{q}}^{\,{\bar{b}}} \nonumber \\=: & {} \textsf{E}\bar{\textsf{E}}\Bigl [\, \tilde{G}(\xi _1,\xi _2)\, \,e^{\,-\,i\sqrt{u}\,\int _0^t[\,b\,d\nu _{12,s}\,+ \,{\bar{b}}\,d\nu _{21,s}\,] }\,\Bigr ] \,q^b\,{\bar{q}}^{\,{\bar{b}}} \end{aligned}$$
(5.85)

where we abbreviated the quantity

$$\begin{aligned} \tilde{G}(\xi _1,\xi _2):= & {} \bigl [\,G(n_{1,t},n_{2,t})\,e^{\,-\,i2u(b-{\bar{b}})\int _0^t p(n_s)\,n_{12,s}\,ds }\,\bigr ](\xi _1,\xi _2) \end{aligned}$$
(5.86)

which depends only on the \(\xi \)-variables, but is independent of the \(\eta \)-variables which only show up in the last exponential in (5.85). This means that as in chapter 4 the \(\eta \)-integrals can be performed and give \(\delta \)-functions for the \(\xi \)-variables. With Fresnel BMs given by, for \(j\in \{1,2\}\) and \(t=t_k=kdt\),

$$\begin{aligned} x_{j,t_k}= & {} \sqrt{dt}\,{\sum }_{\ell =1}^k \phi _{j,\ell } \nonumber \\ y_{j,t_k}= & {} \sqrt{dt}\,{\sum }_{\ell =1}^k \theta _{j,\ell } \end{aligned}$$
(5.87)

the Fresnel measure is

$$\begin{aligned} dFd\bar{F}= & {} \mathop \Pi _{\ell =1}^k e^{\,i\,{\sum }_{j=1}^2{\phi _{j,\ell }^2-\theta _{j,\ell }^2\over 2}}\; {d^2\phi _\ell \,d^2\theta _\ell \over (2\pi )^2} \end{aligned}$$
(5.88)

We write again

$$\begin{aligned} \xi _{j,t_k}\;\;=\;\; {x_{j,t_k}-y_{j,t_k}\over \sqrt{2}} \;\;=\;\;\sqrt{dt}\,{\sum }_{\ell =1}^k {\phi _{j,\ell }-\theta _{j,\ell } \over \sqrt{2}}=: & {} \sqrt{dt}\,{\sum }_{\ell =1}^k \alpha _{j,\ell } \nonumber \\ \eta _{j,t_k}\;\;=\;\; {x_{j,t_k}+y_{j,t_k}\over \sqrt{2}} \;\;=\;\;\sqrt{dt}\,{\sum }_{\ell =1}^k {\phi _{j,\ell }+\theta _{j,\ell } \over \sqrt{2}}=: & {} \sqrt{dt}\,{\sum }_{\ell =1}^k \beta _{j,\ell } \end{aligned}$$
(5.89)

such that the Fresnel measure becomes

$$\begin{aligned} dFd\bar{F}(\alpha ,\beta )= & {} \mathop \Pi _{\ell =1}^k e^{\,i\,(\alpha _{1,\ell }\beta _{1,\ell }+\alpha _{2,\ell }\beta _{2,\ell }) }\; {d^2\alpha _\ell \,d^2\beta _\ell \over (2\pi )^2} \end{aligned}$$
(5.90)

We have to calculate

$$\begin{aligned} { \textsf{E}\bar{\textsf{E}}\Bigl [\, \tilde{G}(\xi _1,\xi _2)\, \,e^{\,-\,i\sqrt{u}\,\int _0^t[\,b\,d\nu _{12,s}\,+ \,{\bar{b}}\,d\nu _{21,s}\,] }\,\Bigr ] \;\;=\;\; } \\{} & {} \int _{{\mathbb R}^{4k}}\, \tilde{G}(\xi _1,\xi _2) \; \exp \bigl \{\,-\,i\sqrt{u}\,b\int _0^t d\nu _{12} \,-\,i\sqrt{u}\,{\bar{b}}\int _0^t d\nu _{21}\, \bigr \} \; dFd\bar{F}(\alpha ,\beta ) \nonumber \end{aligned}$$
(5.91)

with the discrete time expressions

$$\begin{aligned} \int _0^t d\nu _{12}= & {} \int _0^t (\, d\xi _1+d\xi _2 +d\eta _1-d\eta _2\, ) \;\;=\;\; \sqrt{dt}\, {\sum }_{\ell =1}^k \bigl (\, \alpha _{1,\ell }+\alpha _{2,\ell }+\beta _{1,\ell }-\beta _{2,\ell }\,\bigr ) \nonumber \\ \int _0^t d\nu _{21}= & {} \int _0^t (\, d\xi _1+d\xi _2 -d\eta _1+d\eta _2\, ) \;\;=\;\; \sqrt{dt}\, {\sum }_{\ell =1}^k \bigl (\, \alpha _{1,\ell }+\alpha _{2,\ell }-\beta _{1,\ell }+\beta _{2,\ell }\,\bigr ) \ \ \ \ \ \end{aligned}$$
(5.92)

The \(\beta \)-integrals can be performed and produce \(\delta \)-functions:

$$\begin{aligned} { \int _{{\mathbb R}^{2k}} \exp \Bigl \{\,-\,i\sqrt{udt}\, {\sum }_{\ell =1}^k (b-{\bar{b}}) [\beta _{1,m}-\beta _{2,m}]\,\Bigr \} } \mathop \Pi _{\ell =1}^k e^{\,i\,(\alpha _{1,\ell }\beta _{1,\ell }+\alpha _{2,\ell }\beta _{2,\ell }) }\; { d\beta _{1,\ell }\,d\beta _{2,\ell }\over (2\pi )^2} \nonumber \\ \;\;=\;\;\mathop \Pi _{\ell =1}^k \delta \Bigl (\,\alpha _{1,\ell }-[b-{\bar{b}}]\sqrt{udt}\,\Bigr )\;\delta \Bigl (\,\alpha _{2,\ell }+[b-{\bar{b}}]\sqrt{udt}\,\Bigr )\, \end{aligned}$$
(5.93)

Thus,

$$\begin{aligned} \alpha _{1,\ell }= & {} \,+\,(b-{\bar{b}})\sqrt{udt} \nonumber \\ \alpha _{2,\ell }= & {} \,-\,(b-{\bar{b}})\sqrt{udt} \end{aligned}$$
(5.94)

such that \(n_{1,t}\) and \(n_{2,t}\) become, with \(t=t_k=kdt\),

$$\begin{aligned} n_{1,t}= & {} n_1\,e^{\,-\,i\sqrt{4udt}{\sum }_{\ell =1}^k \alpha _{1,\ell } } \;\;=\;\; n_1\,e^{\,-\,i\,2u(b-{\bar{b}})\,t} \nonumber \\ n_{2,t}= & {} n_2\,e^{\,-\,i\sqrt{4udt}{\sum }_{\ell =1}^k \alpha _{2,\ell } } \;\;=\;\; n_2\,e^{\,+\,i\,2u(b-{\bar{b}})\,t} \end{aligned}$$
(5.95)

and therefore

$$\begin{aligned} { \,-\,i2u(b-{\bar{b}})\int _0^t p(n_s)\,n_{12,s}\,ds} \nonumber \\= & {} \,-\,i2u(b-{\bar{b}})\int _0^t p(n_1\,e^{\,-\,i\,2u(b-{\bar{b}})\,s}+n_2\,e^{\,+\,i\,2u(b-{\bar{b}})\,s})\,(n_1\,e^{\,-\,i\,2u(b-{\bar{b}})\,s}-n_2\,e^{\,+\,i\,2u(b-{\bar{b}})\,s})\,ds \nonumber \\= & {} \,+\,\int _0^t p(n_1\,e^{\,-\,i\,2u(b-{\bar{b}})\,s}+n_2\,e^{\,+\,i\,2u(b-{\bar{b}})\,s})\,{d\over ds}(n_1\,e^{\,-\,i\,2u(b-{\bar{b}})\,s}+n_2\,e^{\,+\,i\,2u(b-{\bar{b}})\,s})\,ds \nonumber \\= & {} \,+\,\int _0^t {d\over ds}\bigl [\,\log P(n_1\,e^{\,-\,i\,2u(b-{\bar{b}})\,s}+n_2\,e^{\,+\,i\,2u(b-{\bar{b}})\,s})\,\bigr ]\,ds \nonumber \\= & {} \log P(n_1\,e^{\,-\,i\,2u(b-{\bar{b}})\,t}+n_2\,e^{\,+\,i\,2u(b-{\bar{b}})\,t}) \;-\; \log P(n_1+n_2) \end{aligned}$$
(5.96)

Hence we arrive at

$$\begin{aligned} { \textsf{E}\bar{\textsf{E}}\Bigl [\, G(n_{1,t},n_{2,t})\,e^{\,-\,i2u(b-{\bar{b}})\int _0^t p(n_s)\,n_{12,s}\,ds } \,e^{ \,-\,i\sqrt{u}\,\int _0^t[\,b\,d\nu _{12,s}\,+ \,{\bar{b}}\,d\nu _{21,s}\,] } \,\Bigr ] \;\;=\;\;} \\{} & {} G\bigl (\,e^{-i2u t\,(b-{\bar{b}})}n_1 \,, \, e^{+i2u t\,(b-{\bar{b}})}n_2\,\bigr ) \;\times \; {P(n_1\,e^{\,-\,i\,2u(b-{\bar{b}})\,t}\,+\,n_2\,e^{\,+\,i\,2u(b-{\bar{b}})\,t}) \over P(n_1+n_2) } \nonumber \end{aligned}$$
(5.97)

and this proves part (b) of the theorem. \(\square \)

5.3 Collapse and Revivals

For small u, the two site Bose-Hubbard model shows the intriguing phenomenom of collapse and revivals. In two very beautiful papers, Fishman and Veksler [14] and Bakman, Fishman and Veksler [15] gave a very precise quantitative description of this phenomenom not only for the two site Bose-Hubbard model, but also, to emphasize the general mechanism, for a quantum mechanical oscillator with a small anharmonic perturbation. The main technical tool there was a careful semiclassical analysis of the energy spectrum. Lena Simon and Walter Strunz also used semiclassical methods in their article [16].

Here in our setting we have SDEs and ODEs and of course we want to use them in order to demonstrate the phenomenom. In this paper, we do not aim at the most sophisticated version of doing that, solving this problem would basically mean to solve the quantum mechanical many body problem, but here we just want to give a ‘proof of concept’, namely, to show that the formalism is able to do that at all. To this end, recall the exact equations (5.20) and (5.21) which in the coherent state case read as follows:

$$\begin{aligned} \langle n_t\rangle= & {} N \;\;\;\; \forall t \nonumber \\ {d\over dt}\langle n_{12,t}\rangle= & {} \;+\;i\,2\varepsilon \,\bigl (\,\langle q_t\rangle \,-\,\langle {\bar{q}}_t\rangle \,\bigr ) \nonumber \\ {d\over dt}\langle q_t\rangle= & {} \;+\;i\,\varepsilon \,\langle n_{12,t}\rangle \;-\; i\,2u\,\langle \,n_{12,t}\,q_t\,\rangle \nonumber \\ {d\over dt}\langle {\bar{q}}_t\rangle= & {} \;-\;i\,\varepsilon \,\langle n_{12,t}\rangle \;+\; i\,2u\,\langle \,n_{12,t}\,{\bar{q}}_t\,\rangle \end{aligned}$$
(5.98)

If we would simply factorize \(\,\langle n_{12}\,q\rangle \approx \langle n_{12}\rangle \langle q\rangle \,\), we would recover the mathematical pendulum which does not have collapse and revivals. From part (b) of Theorem 8 of the previous section, we have

$$\begin{aligned} \langle \,n_{1,t}\,q_t\,\rangle _{u}\;\;:=\;\;e^{-it\mathcal{L}_u}\{\, n_{1}q\,\}= & {} \,e^{-i2u t\,}n_1\;\times \; { P(\,e^{-i2u t}n_1 \, + \, e^{+i2u t}n_2\,) \over P(n_1+n_2)}\;q \nonumber \\ \langle \,n_{2,t}\,q_t\,\rangle _{u}\;\;:=\;\;e^{-it\mathcal{L}_u}\{\, n_{2}q\,\}= & {} \,e^{+i2u t\,}n_2\;\times \; { P(\,e^{-i2u t}n_1 \, + \, e^{+i2u t}n_2\,) \over P(n_1+n_2)}\;q \end{aligned}$$
(5.99)

or, since we are considering the coherent state case with \(P(x)=e^x\),

$$\begin{aligned} \langle \,n_{1,t}\,q_t\,\rangle _{u}= & {} \,e^{-i2u t\,}n_1\;\times \; \exp \bigl \{\,(e^{-i2u t}-1)n_1 \, + \, (e^{+i2u t}-1)n_2\,\bigr \} \;\times \;q \nonumber \\ \langle \,n_{2,t}\,q_t\,\rangle _{u}= & {} \,e^{+i2u t\,}n_2\;\times \; \exp \bigl \{\,(e^{-i2u t}-1)n_1 \, + \, (e^{+i2u t}-1)n_2\,\bigr \} \;\times \;q \end{aligned}$$
(5.100)

We also have

$$\begin{aligned} \langle n_{1,t}\rangle _{u}\;\;=\;\;e^{-it\mathcal{L}_u} \,n_{1}= & {} \, n_1 \nonumber \\ \langle n_{2,t}\rangle _{u}\;\;=\;\;e^{-it\mathcal{L}_u} \,n_{2}= & {} \, n_2 \end{aligned}$$
(5.101)

and

$$\begin{aligned} \langle q_t\rangle _{u}\;\;=\;\;e^{-it\mathcal{L}_u}\,q= & {} \,\exp \bigl \{\,(e^{-i2u t}-1)n_1 \, + \, (e^{+i2u t}-1)n_2\,\bigr \} \;\times \;q \end{aligned}$$
(5.102)

Thus, for the dynamics under \(\varepsilon =0\), we can write

$$\begin{aligned} {\langle \,n_{1,t}\,q_t\,\rangle _{u} \over \langle n_{1,t}\rangle _{u}\,\langle q_t\rangle _{u}}= & {} \,e^{\,-\,i2u t\,} \nonumber \\ {\langle \,n_{2,t}\,q_t\,\rangle _{u} \over \langle n_{2,t}\rangle _{u}\,\langle q_t\rangle _{u}}= & {} \,e^{\,+\,i2u t\,} \end{aligned}$$
(5.103)

Now consider the dynamics under \(u=0\). Since \(e^{-it\mathcal{L}_\varepsilon }\) simply rotates the argument when applied to an arbitrary function F,

$$\begin{aligned} \bigl \langle \,F(n_{1,t},n_{2,t},q_t,{\bar{q}}_t)\,\bigr \rangle _\varepsilon \;\;:=\;\;(e^{-it\mathcal{L}_\varepsilon }F)(n_1,n_2,q,{\bar{q}})= & {} F\bigl (\, R_t\,(n_1,n_2,q,{\bar{q}})^T \,\bigr ) \end{aligned}$$
(5.104)

the action of \(e^{-it\mathcal{L}_\varepsilon }\) factorizes when applied to an arbitrary product,

$$\begin{aligned} e^{-it\mathcal{L}_\varepsilon }\,(FG)= & {} e^{-it\mathcal{L}_\varepsilon }\,F \;\times \; e^{-it\mathcal{L}_\varepsilon }\,G \end{aligned}$$
(5.105)

Thus, under \(u=0\),

$$\begin{aligned} {\langle \,n_{1,t}\,q_t\,\rangle _{\varepsilon } \over \langle n_{1,t}\rangle _{\varepsilon }\,\langle q_t\rangle _{\varepsilon }}= & {} \;1 \nonumber \\ {\langle \,n_{2,t}\,q_t\,\rangle _{\varepsilon } \over \langle n_{2,t}\rangle _{\varepsilon }\,\langle q_t\rangle _{\varepsilon }}= & {} \;1 \end{aligned}$$
(5.106)

Then, for the full dynamics with both \(\varepsilon \) and u being nonzero, one may try the approximation

$$\begin{aligned} {\langle \,n_{1,t}\,q_t\,\rangle \over \langle n_{1,t}\rangle \,\langle q_t\rangle }\approx & {} \;{1\over 2}\,\Bigl (\,1\,+\,e^{\,-\,i2u t\,}\,\Bigr ) \nonumber \\ {\langle \,n_{2,t}\,q_t\,\rangle \over \langle n_{2,t}\rangle \,\langle q_t\rangle }\approx & {} \;{1\over 2}\,\Bigl (\,1\,+\,e^{\,+\,i2u t\,}\,\Bigr ) \end{aligned}$$
(5.107)

and this in fact generates collapse and revivals. That is, we modify the exact system (5.98) to the following approximate system (recall the abbreviation \(\,n_{12}:=n_1-n_2\,\))

$$\begin{aligned} \langle n_t\rangle= & {} N \;\;\;\; \forall t \\ {d\over dt}\langle n_{12,t}\rangle= & {} \;+\;i\,2\varepsilon \,\bigl (\,\langle q_t\rangle \,-\,\langle {\bar{q}}_t\rangle \,\bigr ) \nonumber \\ {d\over dt}\langle q_t\rangle\approx & {} \;+\;i\,\varepsilon \,\langle n_{12,t}\rangle \;-\; i\,u\,\bigl \{\,(1+e^{-i2ut})\,\langle n_{1,t}\rangle \langle q_t\rangle \,-\,(1+e^{+i2ut})\,\langle n_{2,t}\rangle \langle q_t\rangle \,\bigr \} \nonumber \\ {d\over dt}\langle {\bar{q}}_t\rangle\approx & {} \;-\;i\,\varepsilon \,\langle n_{12,t}\rangle \;+\; i\,u\,\bigl \{\,(1+e^{+i2ut})\,\langle n_{1,t}\rangle \langle {\bar{q}}_t\rangle \,-\,(1+e^{-i2ut})\,\langle n_{2,t}\rangle \langle {\bar{q}}_t\rangle \,\bigr \} \nonumber \end{aligned}$$
(5.108)

which then reduces to the following two equations (since \(\,\langle n_t\rangle =\langle n_{1,t}+n_{2,t}\rangle =N\,\))

$$\begin{aligned} {d\over dt}\langle n_{12,t}\rangle= & {} \;-\;4\varepsilon \;\textrm{Im}\langle q_t\rangle \\ {d\over dt}\langle q_t\rangle\approx & {} \;+\;i\,\varepsilon \,\langle n_{12,t}\rangle \;-\; i\,u\,(1+\cos 2ut)\,\langle n_{12,t}\rangle \langle q_t\rangle \,-\,uN\sin 2ut \,\langle q_t\rangle \nonumber \end{aligned}$$
(5.109)

Before we look at the numerical results, let’s make some quick analytical considerations. Since from now on we are purely in the ODE framework, let’s omit the angular brackets and summarize the system as follows:

$$\begin{aligned} \dot{n}_{12,t}= & {} \;-\;4\varepsilon \,\textrm{Im}\, q_t \nonumber \\ \dot{q}_t= & {} \;+\;i\,\varepsilon \,n_{12,t}\;-\; i\,u\,(1+\cos 2{\tilde{u}}t)\,n_{12,t}\, q_t\,-\,uN\sin 2{\tilde{u}}t \, q_t \end{aligned}$$
(5.110)

Here the \({\tilde{u}}\) on the right hand side of (5.110) is actually a u, we put this in since for \({\tilde{u}}=0\) this reduces to the system (5.27),(5.28) of Section 5.1 where we have shown that this is actually the mathematical pendulum. Thus, by switching the \({\tilde{u}}\) from 0 to u we can interpolate between the mathematical pendulum without collapse and revivals and the actual case under consideration. We write

$$\begin{aligned} \ddot{n}_{12,t}= & {} \;-\;4\varepsilon \,\textrm{Im}\, \dot{q}_t \nonumber \\= & {} \;-\;4\varepsilon ^2\,n_{12,t} \;+\;4\varepsilon \,u\,(1+\cos 2{\tilde{u}}t)\,n_{12,t}\, \textrm{Re}\, q_t\;+\; 4\varepsilon uN\sin 2{\tilde{u}}t \, \textrm{Im}\, q_t \nonumber \\= & {} \;-\;4\varepsilon ^2\,n_{12,t} \;+\;4\varepsilon \,u\,(1+\cos 2{\tilde{u}}t)\,n_{12,t}\, \textrm{Re}\, q_t\;-\; uN\sin 2{\tilde{u}}t \, \dot{n}_{12,t} \end{aligned}$$
(5.111)

or

$$\begin{aligned} \ddot{n}_{12,t}\;+\;uN\sin 2{\tilde{u}}t \, \dot{n}_{12,t} \;+\;4\varepsilon ^2\,n_{12,t}= & {} \;+\;4\varepsilon \,u\,(1+\cos 2{\tilde{u}}t)\,n_{12,t}\textrm{Re}\, q_t \end{aligned}$$
(5.112)

For \({\tilde{u}}=0\), the case without collapse and revivals, this reduces to

$$\begin{aligned} \ddot{n}_{12,t}\;+\;0 \;+\;4\varepsilon ^2\,n_{12,t}= & {} \;+\;4\varepsilon \,u\,(1+1)\,n_{12,t}\textrm{Re}\, q_t \end{aligned}$$
(5.113)

That is, the right hand side of (5.112) is basically responsible for the difference between a harmonic pendulum and the mathematical pendulum and it generates the effect of self trapping or rollovers for \(g>2\varepsilon \). Since collapse and revivals do already show up for very small g where surely the harmonic approximation should be valid, we should be able to see them already in the following equation

$$\begin{aligned} \ddot{n}_{12,t}\;+\;uN\sin 2{\tilde{u}}t \, \dot{n}_{12,t} \;+\;4\varepsilon ^2\,n_{12,t}= & {} 0 \end{aligned}$$
(5.114)

This is simply a harmonic oscillator with a time dependent friction

$$\begin{aligned} \gamma _t:= & {} uN\sin 2{\tilde{u}}t \end{aligned}$$
(5.115)

which can be transformed away with the Ansatz

$$\begin{aligned} n_{12,t}= & {} e^{\,-\,{1\over 2}\int _0^t \gamma _s\, ds} \; y_t \end{aligned}$$
(5.116)

which then produces the following equation for \(y_t\),

$$\begin{aligned} \ddot{y}_t \;+\; {\omega }_t^2\, y_t= & {} 0 \end{aligned}$$
(5.117)

with a time dependent frequency

$$\begin{aligned} {\omega }_t^2= & {} 4\varepsilon ^2\,-\,{\gamma _t^2\over 4} \,-\,{\dot{\gamma }_t\over 2} \;\;=\;\; 4\varepsilon ^2\,-\,{(uN)^2\sin ^2 2{\tilde{u}}t\over 4} \,-\,{\tilde{u}}\,uN\cos 2{\tilde{u}}t \end{aligned}$$
(5.118)

For small \(g=uN\), one may put this roughly to \(4\varepsilon ^2\) such that \(y_t\) is identical to the exact \(u=0\) solution which is

$$\begin{aligned} y_t\approx & {} N\cos 2\varepsilon t \end{aligned}$$
(5.119)

Thus, collapse and revivals arise from the damping factor

$$\begin{aligned} e^{\,-\,{1\over 2}\int _0^t \gamma _s\, ds}= & {} \exp \bigl \{\,-\,{uN\over 2}\int _0^t \sin 2{\tilde{u}}s\, ds \, \bigr \} \;\; \buildrel {\tilde{u}}=u\over = \;\; \exp \bigl \{\,-\,{N\over 4}\,(1-\cos 2ut)\, \bigr \} \end{aligned}$$
(5.120)

and we arrive at the approximate small g solution

$$\begin{aligned} n_{12,t}\approx & {} \exp \bigl \{\,-\,{N\over 4}\,(1-\cos 2ut)\, \bigr \} \;\times \; N\cos 2\varepsilon t \end{aligned}$$
(5.121)

In the following Fig. 3 and 4, the quantity \(n_{12,t}/N\) obtained from exact diagonalization, in black, is plotted together with the analytical collapse factor (5.120), in red. We chose \(N=50\), \(\varepsilon =1\) and g as displayed on the plots and all 50 particles were put onto lattice site 1 at time \(t=0\,\):

However, now we have to remark that the exact diagonalization numbers were produced with a number state, not with a coherent state. While in the large N limit the dynamics of coherent states and number states are identical, this no longer holds in the revival region after the first collapse has occured. In the large N limit, this region moves to infinity and hence is not visible there. For example, if we put \(g=0.02\) and choose a time horizon of \(T=10000\,\pi \) instead of \(T=1000\,\pi \) as above, a comparison of number state and coherent state dynamics looks as displayed in Fig. 5.

This clearly demonstrates that the problem is subtle and a more careful analysis is required. For example, in the approximation (5.107) we could have equally well have said, we use \(e^{-i2ut}\) instead of an average \((1+e^{-i2ut})/2\), and then this would have had the effect that the analytical collapse factor in (5.120) would have come with an N/2 in the exponent instead of an N/4 and the revival blobs were too small. So, this really should be considered as some kind of a ‘teaser’, some kind of a motivational argument, but nothing more. A proper and systematic treatment of the diffusive part is basically equivalent to solving the quantum mechanical many body problem and this still needs to be developed. But to do so, we believe indeed that the formalism presented in this paper is a very useful tool.

Fig. 3
figure 3

Exact diagonalization vs. analytical collapse factor \( \exp \bigl \{\,-\,{N\over 4}\,(1-\cos 2ut)\,\bigr \}\)

Fig. 4
figure 4

Exact diagonalization vs. analytical collapse factor \(\exp \{-\frac{N}{4}(1-\cos 2ut)\}\), other g’s

Fig. 5
figure 5

Coherent state vs. number state dynamics

Let us remark that the qualitative behaviour seen in Fig. 5 below, number states have twice the number of revivals than coherent states, actually can be made plausible by comparing their respective collapse factors: for coherent states, there is basically a factor of \(\,\exp \{-N[1-\cos (2ut)]\}\,\) (ignoring the width of the revival blobs for the moment) while for number states there is a factor \(\,[\cos (2ut)]^N\,\). In the first case, we have revivals at \(\,\cos (2ut)=+1\,\) while in the second case, we have revivals at \(\,\cos (2ut)=+1\,\) but also for \(\,\cos (2ut)=-1\,\), so just twice as much, as seen in the numerical simulation.

We also implemented the full approximate ODE system (5.109) and compared with exact diagonalization numbers, again using a number state, not a coherent state. We used a fourth order Runge-Kutta method and separated off the collapse factor to obtain a stable implementation (\(N=50\) and \(\varepsilon =1\) as above, g fixed to 0.1 and zoomed in on different time windows), the exact diagonalization numbers are in black and the ODE solution is in red in the following Figs. 6:

Fig. 6
figure 6

Exact diagonalization (black) and ODE solution (red) on different time windows

6 Summary

The paper has demonstrated that the formalism of stochastic calculus is very useful to address the dynamics of the Bose-Hubbard model. The fact that in the large N limit the exact quantum dynamics can be obtained from an ODE system, the time dependent discrete GP equation, has been derived in a conceptually very pure and clean and transparent way. For finite N, the dynamics is given by the SDE systems of Theorems 4 and 5 of chapter 2 and the diffusive parts of those systems vanish in the large N limit. More generally, the paper provides a technique to obtain GP-like mean field equations for an arbitrary given initial state, in arbitrary dimension and for an arbitrary hopping matrix. For the two site Bose-Hubbard model, the diffusive part has been taken into account with an approximation and collapse and revivals could be reproduced, numerically and also through an analytic calculation. A proper systematic treatment of the diffusive part is still missing and needs to be developed. It has also been shown that density matrix elements can be obtained from various exact parabolic second order PDEs.

7 Additional Remarks

By the end of 2016, the mathematics department of Hochschule RheinMain joined the Faculty of Engineering and the author was asked by Klaus Michael Indlekofer from Electrical Engineering whether there would be some interest in joining a project on quantum dynamics. After 9 years as a financial engineer at a bank, the author found that this would be a good opportunity to reenter the field and it didn’t took long until it was realized that the Hubbard model is more relevant than ever due to some major experimental breakthroughs in the ultracold atoms area [17,18,19,20,21]. Working purely on the theoretical side, we can only humbly take notice of what is doable there [22].

First attention then was drawn to phase space methods and the truncated Wigner approximation because of the very attractive idea to get the quantum dynamics from suitably weighted ODE trajectories. In particular, the beautiful papers of Polkovnikov [23,24,25], Polkovnikov, Sachdev and Girvin [26] and Davidson, Sels and Polkovnikov [27] served as a major motivation and inspiration for the current work.

With a theoretical and practical background in stochastic calculus from 9 years of option pricing, then it was natural to take a closer look to the long history of stochastic methods applied to the quantum many body problem [28, 29]. In particular, the formalism of the Husimi Q-Function and the Positve P-Representation [30,31,32,33,34] was considered more closely and this, combined with the background of the author [35, 36], then lead to the approach which is taken in this paper.

Nowadays nearly taken for granted, but the almost unlimited and instantaneous access to the science knowledge of the planet and the people who provide it also has been critical for the completion of this work. There have been numerous papers, the majority of them probably not being cited here, where just a particular item was looked up and then the conclusion was, okay, for our purposes this does not lead in the right direction. Those references may not seem directly relevant to the now final version of this paper, but they have been critical in order to get there. In this class fall for example references for BCH like formulas and time ordered exponentials [37,38,39,40] ([39] derives very interesting formulae which are in the same spirit but more general than a formula derived by the author in chapter 10 of [36]), Carleman Linearization Technique and Kroenecker products of matrices (Kowalski and Steeb [41] wrote a very beautiful book on that) or references which relate to the author’s attempt to evaluate the Fresnel integrals directly.

With the presented formalism, the paper opens up the possibility to address a range of very interesting topics like fermionic models or thermodynamic quantities with an \(e^{-\beta H}\) in it instead of an \(e^{-itH}\), and the author looks very much forward to consider these issues, but with a teaching load of 18 hours per week at a German University of Applied Sciences, research basically has to be restricted to the off-term periods which are March and August and September each year.