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

11institutetext: Division of Liberal Arts and Sciences, Aichi-Gakuin University
11email: [email protected]

The Berezin–Simon quantization for Kähler manifolds and their path integral representations

Hideyasu Yamashita
Abstract

The Berezin–Simon (BS) quantization is a rigorous version of the “operator formalism” of quantization procedure. The goal of the paper is to present a rigorous real-time (not imaginary-time) path-integral formalism corresponding to the BS operator formalism of quantization; Here we consider the classical systems whose phase space MM is a (possibly non-compact) Kähler manifold which satisfies some conditions, with a Hamiltonian H:MH:M\to\mathbb{R}. For technical reasons, we consider only the cases where HH is smooth and bounded. We use Güneysu’s extended version of the Feynman–Kac theorem to formulate the path-integral formula.

1 Introduction

In this paper, we set the “classical phase space” 𝐌{\bf M} to be a (possibly non-compact) Kählerian manifold which is a submanifold of a (possibly infinite-dimensional) complex projective space \mathbb{P}\mathcal{H}, where \mathcal{H} is a complex Hilbert space. The phase space 𝐌{\bf M} admits a quantization procedure, which we called the Glauber-Sudarshan-type quantization in [15]; However instead we will call it the Berezin–Simon (BS) quantization in this paper, since we follow the formulation of the quantization procedure given in Simon [12], which is based on Berezin’s works; See [12] and references therein. A BS quantization is an “operator formalism” of quantization procedure. The goal of the paper is to present a path-integral representation of the BS quantization; Roughly, we present a rigorous path-integral formalism corresponding to the BS operator formalism of quantization.

The previous paper [15] had a similar goal, but it was very restricted in that we confined ourselves to the cases where the phase space 𝐌{\bf M} is a compact homogeneous space, which is a submanifold of a projective space \mathbb{P}\mathcal{H} where \mathcal{H} is a finite-dimensional Hilbert space. Thus this paper will be viewed as a considerable extension of [15].

Our main mathematical tool for the path-integral formulation is the Feynman-Kac formula on a vector bundle over a Riemannian manifold given by [5], together with the Bochner-Kodaira-Nakano identity for Kählerian manifolds. A Feynman-Kac formula itself is seen as a mathematical justification of the imaginary-time path integral method in quantum physics, but we devise a method to use it for real-time path integrals.

Roughly speaking, this paper is situated in the following context of past rigorous studies on path integrals.

Feynman’s original idea [4] is to represent the time evolution of a quantum system, as well as the expectation values of observables in it, by an integral on the space of paths on the configuration space of the system. As is well known, if we consider the “imaginary time” evolution instead of real time evolution, so-called the Wick rotation, a large part of the idea can be made rigorous by the Feynman–Kac theorem and its generalizations, and this “imaginary time ++ Feynman–Kac” approach is the most successful one. However, note that in the imaginary-time approaches, it is difficult to deal with time-dependent Hamiltonians, as well as non-unitary time evolutions occurring in open systems. This implies that it is hard to apply the imaginary-time methods to e.g. the theories of quantum information/probability, where time-dependent Hamiltonians and non-unitary time evolutions (e.g. decoherences) frequently occur.

On the other hand, the notion on configuration-space path integrals are believed to be derived from more general notion of phase-space path integrals. In some sense, the latter ones may be more fundamental if we consider a path integral as a procedure of quantization of a classical system; The main stream of the rigorous studies of quantization (e.g. the theories of geometric/deformation quantization) are formulated on phase spaces. Unlike imaginary-time configuration-space path integrals, little is known about the rigorous justification of general phase-space path integrals (in real or imaginary time).

However in 1985, Daubechies and Klauder [3] gave an important rigorous result on coherent-state path integrals, which can be viewed as a sort of phase-space path integral formula, representing real-time evolution for some class of Hamiltonians, in terms of Brownian motions and stochastic integrals. Yamashita [14] studied phase-space path integrals in a similar idea but for other class of Hamiltonians, and with an emphasis on geometric meaning of them. In these results, mainly the phase spaces are assumed to be flat, i.e., 𝐌2nn{\bf M}\cong\mathbb{R}^{2n}\cong\mathbb{C}^{n}. Yamashita [15] is seen as an attempt to apply such methods on some sort of (non-flat) compact phase spaces, which arise in irreducible unitary representation of semisimple Lie groups, e.g., SU(n){\rm SU}(n), SO(n){\rm SO}(n), Sp(n){\rm Sp}(n), etc. Then we are given a question whether these works [14, 15] can be unified and extended for more general phase spaces, or not. This paper is intended to be an affirmative answer to this question.

However note that we consider only the bounded Hamiltonians in this paper, hence not all of the results of [3] is contained in our result. Since this boundedness assumption is quite unsatisfactory for applications to realistic physical systems, we are required to loosen this assumption, but the treatment of general unbounded Hamiltonians appears to be extremely difficult. A hopeful approach will be to examine some moderate assumptions such as that the classical Hamiltonian H(x)H(x) is bounded from below and increases as H(x)|x|2H(x)\sim|x|^{2}; Another hopeful approach will be to consider a “solvable” (or “algebraically tractable”) set of Hamiltonians which are generators of (representations of) a Lie group, e.g. symplectic groups, Poincare groups, etc. The latter approach will be related to the construction of unitary representations of a Lie group in terms of orbit method/geometric quantization [6, 13].

2 Projective representations of BS quantizations

2.1 BS quantization

Let \mathcal{H} be a complex Hilbert space, and \mathbb{P}\mathcal{H} denote the set of orthogonal projections onto one-dimensional subspaces of \mathcal{H}, that is,

:={|vv|:v,v=1}.\mathbb{P}\mathcal{H}:=\left\{|v\rangle\langle v|:\ v\in\mathcal{H},\ \|v\|=1\right\}.

Let ×:={0}\mathcal{H}^{\times}:=\mathcal{H}\setminus\{0\}, and define 𝐩𝐫:×{\bf pr}:\mathcal{H}^{\times}\to\mathbb{P}\mathcal{H} to be the map from v×v\in\mathcal{H}^{\times} to the orthogonal projection onto v\mathbb{C}v, i.e.,

𝐩𝐫(v):=|vv|v2,v×.{\bf pr}(v):=\frac{|v\rangle\langle v|}{\left\|v\right\|^{2}},\qquad v\in\mathcal{H}^{\times}.

Let 𝐌{\bf M} be a subset of \mathbb{P}\mathcal{H} with the measure μ\mu. The measure space (𝐌,μ)({\bf M},\mu) is called a family of coherent states on \mathcal{H} if

𝐌𝐩dμ(𝐩)=I.\int_{{\bf M}}\mathbf{p}\,{\rm d}\mu(\mathbf{p})=I.

Let Coh(){\rm Coh}(\mathcal{H}) be the set of families of coherent states on \mathcal{H}. For a function f:f:\mathbb{P}\mathcal{H}\to\mathbb{C}, let

𝒬(f):=𝐌f(𝐩)𝐩dμ(𝐩),\mathcal{Q}(f):=\int_{{\bf M}}f(\mathbf{p})\mathbf{p}\,{\rm d}\mu(\mathbf{p}),

if the integral exists. In this paper, we call the operation f𝒬(f)f\mapsto\mathcal{Q}(f) the BS quantization on (𝐌,μ)({\bf M},\mu).

Let

𝐩𝐫1(𝐌):={v×|𝐩𝐫(v)𝐌}=𝐩𝐌ran(𝐩){0}{\bf pr}^{-1}({\bf M}):=\left\{v\in\mathcal{H}^{\times}|{\bf pr}(v)\in{\bf M}\right\}=\bigcup_{\mathbf{p}\in{\bf M}}\mathrm{ran}(\mathbf{p})\setminus\{0\}
𝕊():={v|v=1},𝕊(𝐌):=𝕊()𝐩𝐫1(𝐌)\mathbb{S}(\mathcal{H}):=\left\{v\in\mathcal{H}|\left\|v\right\|=1\right\},\quad\mathbb{S}({\bf M}):=\mathbb{S}(\mathcal{H})\cap{\bf pr}^{-1}({\bf M})

The projection 𝐩𝐫|𝕊(𝐌):𝕊(𝐌)𝐌{\bf pr}|_{\mathbb{S}({\bf M})}:\mathbb{S}({\bf M})\to{\bf M} defines a U(1)U(1)-bundle over 𝐌{\bf M}. (Precisely, the term “bundle” should be used only when 𝐌{\bf M} is a (smooth) manifold.) Let μ𝕊\mu_{\mathbb{S}} be the U(1)U(1)-invariant measure on 𝕊(𝐌)\mathbb{S}({\bf M}) determined by

μ(E)=μ𝕊(𝐩𝐫𝕊1(E)),E𝐌,measurable\mu(E)=\mu_{\mathbb{S}}({\bf pr}_{\mathbb{S}}^{-1}(E)),\qquad\forall E\subset{\bf M},\ \text{measurable}

where 𝐩𝐫𝕊:=𝐩𝐫|𝕊(𝐌){\bf pr}_{\mathbb{S}}:={\bf pr}|_{\mathbb{S}({\bf M})}. For measurable functions f1,f2:𝕊(𝐌)f_{1},f_{2}:\mathbb{S}({\bf M})\to\mathbb{C} (or f1,f2:𝐩𝐫1(𝐌)f_{1},f_{2}:{\bf pr}^{-1}({\bf M})\to\mathbb{C}), let

f1|f2𝕊:=𝕊(𝐌)f1(v)¯f2(v)dμ𝕊(v),\left\langle f_{1}|f_{2}\right\rangle_{\mathbb{S}}:=\int_{\mathbb{S}({\bf M})}\overline{f_{1}(v)}f_{2}(v){\rm d}\mu_{\mathbb{S}}(v),

if the integral exists.

Let \mathcal{H}^{*} denote the dual of \mathcal{H}. For any uu\in\mathcal{H}, define uu^{*}\in\mathcal{H}^{*} by

u(v):=u|v,v.u^{*}(v):=\left\langle u|v\right\rangle,\qquad v\in\mathcal{H}.

We denote u|𝕊(𝐌)u^{*}|_{\mathbb{S}({\bf M})} simply by uu^{*}. Then we find that

u1|u2𝕊=u2|u1,u1,u2.\left\langle u_{1}^{*}|u_{2}^{*}\right\rangle_{\mathbb{S}}=\left\langle u_{2}|u_{1}\right\rangle,\qquad\forall u_{1},u_{2}\in\mathcal{H}.

The inner product |𝕊\left\langle\cdot|\cdot\right\rangle_{\mathbb{S}} defines the Hilbert space L2(𝕊(𝐌))L2(𝕊(𝐌),μ𝕊)L^{2}(\mathbb{S}({\bf M}))\equiv L^{2}(\mathbb{S}({\bf M}),\mu_{\mathbb{S}}).

For \ell\in\mathbb{Z}, let

Γ(×)\displaystyle\Gamma_{\ell}(\mathcal{H}^{\times}) :={f:×:f(λv)=λnf(v),λ×,v×},\displaystyle:=\left\{f:\mathcal{H}^{\times}\to\mathbb{C}\,:\,f(\lambda v)=\lambda^{n}f(v),\ \ \forall\lambda\in\mathbb{C}^{\times},\ v\in\mathcal{H}^{\times}\right\},
Γ,𝐌Γ(𝐩𝐫1(𝐌))\displaystyle\Gamma_{\ell,{\bf M}}\equiv\Gamma_{\ell}({\bf pr}^{-1}({\bf M})) :={f:𝐩𝐫1(𝐌):f(λv)=λf(v),λ×,v𝐩𝐫1(𝐌)},\displaystyle:=\left\{f:{\bf pr}^{-1}({\bf M})\to\mathbb{C}\,:\,f(\lambda v)=\lambda^{\ell}f(v),\ \ \forall\lambda\in\mathbb{C}^{\times},\ v\in{\bf pr}^{-1}({\bf M})\right\},
Γ,𝐌Lp\displaystyle\Gamma_{\ell,{\bf M}}^{L^{p}} :={fΓ,𝐌:f|𝕊(𝐌)Lp(𝕊(𝐌),μ𝕊)|},1p.\displaystyle:=\left\{f\in\Gamma_{\ell,{\bf M}}:\,f|_{\mathbb{S}({\bf M})}\in L^{p}(\mathbb{S}({\bf M}),\mu_{\mathbb{S}})|\right\},\qquad 1\leq p\leq\infty.

For each \ell\in\mathbb{Z}, Γ,𝐌L2\Gamma_{\ell,{\bf M}}^{L^{2}} is a Hilbert space with the inner product |𝕊\left\langle\cdot|\cdot\right\rangle_{\mathbb{S}}, and naturally viewed as a closed subspace of L2(𝕊(𝐌),μ𝕊)L^{2}(\mathbb{S}({\bf M}),\mu_{\mathbb{S}}). In this paper, we deal with the cases where =1,0,1\ell=1,0,-1, and our main concern is the case =1\ell=1.

For each uu\in\mathcal{H}, we see u|×Γ1,𝐌L2u^{*}|_{\mathcal{H}^{\times}}\in\Gamma_{1,{\bf M}}^{L^{2}}. Hence \mathcal{H}^{*} is viewed as a closed subspace of Γ1,𝐌L2\Gamma_{1,{\bf M}}^{L^{2}}.

For F:𝐌F:{\bf M}\to\mathbb{C}, define F~:𝐩𝐫1(𝐌)\tilde{F}:{\bf pr}^{-1}({\bf M})\to\mathbb{C} by

F~(v):=F(𝐩𝐫(v)).\tilde{F}(v):=F({\bf pr}(v)).

Then we see F~Γ0,𝐌\tilde{F}\in\Gamma_{0,{\bf M}}; It follows that 𝐌\mathbb{C}^{{\bf M}} (the space of functions 𝐌{\bf M}\to\mathbb{C}) can be identified with Γ0,𝐌\Gamma_{0,{\bf M}}; We can also identify Lp(𝐌,)L^{p}({\bf M},\mathbb{C}) with Γ0,𝐌Lp\Gamma_{0,{\bf M}}^{L^{p}}.

Let FL(𝐌,)F\in L^{\infty}({\bf M},\mathbb{C}), then FF acts on Γ,𝐌L2\Gamma_{\ell,{\bf M}}^{L^{2}} as a pointwise multiplication operator MFM_{F}:

(MFf)(v):=F~(v)f(v),v𝐩𝐫1(𝐌),fΓ,𝐌L2.\left(M_{F}f\right)(v):=\tilde{F}(v)f(v),\qquad v\in{\bf pr}^{-1}({\bf M}),\ f\in\Gamma_{\ell,{\bf M}}^{L^{2}}.

Let EE_{\mathcal{H}^{*}} be the orthogonal projection from Γ1,𝐌L2\Gamma_{1,{\bf M}}^{L^{2}} onto \mathcal{H}^{*}. We see

E=𝕊(𝐌)𝐩𝐫(v)dμ𝕊(v).E_{\mathcal{H}^{*}}=\int_{\mathbb{S}({\bf M})}{\bf pr}(v^{*}){\rm d}\mu_{\mathbb{S}}(v). (2.1)

where 𝐩𝐫(v){\bf pr}(v^{*}) is the orthogonal projection from Γ1,𝐌L2\Gamma_{1,{\bf M}}^{L^{2}} onto v\mathbb{C}v^{*}. For fΓ1,𝐌L2f\in\Gamma_{1,{\bf M}}^{L^{2}}, v𝐩𝐫1(𝐌)v\in{\bf pr}^{-1}({\bf M}), we have

(Ef)(v)=v|Ef𝕊=v|f𝕊=𝕊(𝐌)x|vf(x)dμ𝕊(x).\left(E_{\mathcal{H}^{*}}f\right)(v)=\left\langle v^{*}|E_{\mathcal{H}^{*}}f\right\rangle_{\mathbb{S}}=\left\langle v^{*}|f\right\rangle_{\mathbb{S}}=\int_{\mathbb{S}({\bf M})}\left\langle x|v\right\rangle f(x){\rm d}\mu_{\mathbb{S}}(x).

That is, the integral kernel E(u,v)E_{\mathcal{H}^{*}}(u,v) of EE_{\mathcal{H}^{*}} is given by

E(v,u)=u|v,u,v𝕊(𝐌).E_{\mathcal{H}^{*}}(v,u)=\left\langle u|v\right\rangle,\qquad u,v\in\mathbb{S}({\bf M}). (2.2)

For FL(𝐌,)F\in L^{\infty}({\bf M},\mathbb{C}), define the operator 𝒬~(F)\tilde{\mathcal{Q}}(F) on Γ1,𝐌L2\Gamma_{1,{\bf M}}^{L^{2}} by

𝒬~(F):=𝕊(𝐌)F~(v)𝐩𝐫(v)dμ𝕊(v).\tilde{\mathcal{Q}}(F):=\int_{\mathbb{S}({\bf M})}\tilde{F}(v){\bf pr}(v^{*}){\rm d}\mu_{\mathbb{S}}(v). (2.3)

We see

𝒬~(F)v=(𝒬(F)v),v|𝒬~(F)u𝕊=u|𝒬(F)v,u,v.\tilde{\mathcal{Q}}(F)v^{*}=\left(\mathcal{Q}(F)^{*}v\right)^{*},\quad\left\langle v^{*}|\tilde{\mathcal{Q}}(F)u^{*}\right\rangle_{\mathbb{S}}=\left\langle u|\mathcal{Q}(F)v\right\rangle,\qquad u,v\in\mathcal{H}.

We call the operation 𝒬~\tilde{\mathcal{Q}} the BS quantization on Γ1,𝐌L2\Gamma_{1,{\bf M}}^{L^{2}} .

The following theorem says that all the information of the BS quantization 𝒬~\tilde{\mathcal{Q}} (or 𝒬\mathcal{Q}) is essentially contained in the projection operator EE_{\mathcal{H}^{*}}.

Theorem 2.1.

For any FL(𝐌,)F\in L^{\infty}({\bf M},\mathbb{C}), we have

𝒬~(F)=EMFE.\tilde{\mathcal{Q}}(F)=E_{\mathcal{H}^{*}}M_{F}E_{\mathcal{H}^{*}}.
Proof.

Let uu\in\mathcal{H} and v×v\in\mathcal{H}^{\times}. Then we have

(EMFEu)(v)\displaystyle\left(E_{\mathcal{H}^{*}}M_{F}E_{\mathcal{H}^{*}}u^{*}\right)(v) =(EMFu)(v)=v|MFu𝕊\displaystyle=\left(E_{\mathcal{H}^{*}}M_{F}u^{*}\right)(v)=\left\langle v^{*}|M_{F}u^{*}\right\rangle_{\mathbb{S}}
=𝕊(𝐌)dμ𝕊(x)v(x)¯F~(x)u(x)\displaystyle=\int_{\mathbb{S}({\bf M})}{\rm d}\mu_{\mathbb{S}}(x)\overline{v^{*}(x)}\tilde{F}(x)u^{*}(x)
=𝕊(𝐌)dμ𝕊(x)v|x¯F~(x)u|x\displaystyle=\int_{\mathbb{S}({\bf M})}{\rm d}\mu_{\mathbb{S}}(x)\overline{\left\langle v|x\right\rangle}\tilde{F}(x)\left\langle u|x\right\rangle
=𝕊(𝐌)dμ𝕊(x)v|x𝕊x|u𝕊F~(x)\displaystyle=\int_{\mathbb{S}({\bf M})}{\rm d}\mu_{\mathbb{S}}(x)\left\langle v^{*}|x^{*}\right\rangle_{\mathbb{S}}\left\langle x^{*}|u^{*}\right\rangle_{\mathbb{S}}\tilde{F}(x)
=v|𝒬~(F)u𝕊=(𝒬~(F)u)(v).\displaystyle=\left\langle v^{*}|\tilde{\mathcal{Q}}(F)u^{*}\right\rangle_{\mathbb{S}}=\left(\tilde{\mathcal{Q}}(F)u^{*}\right)(v).

On the other hand, if fΓ1,𝐌L2f\in\Gamma_{1,{\bf M}}^{L^{2}} is orthogonal to \mathcal{H}^{*},

𝒬~(F)f=0=EMFEf.\tilde{\mathcal{Q}}(F)f=0=E_{\mathcal{H}^{*}}M_{F}E_{\mathcal{H}^{*}}f.

2.2 Some lemmas

Generally, for a bounded operator AA on \mathcal{H}, define the operator A~\tilde{A} on Γ1,𝐌L2\Gamma_{1,{\bf M}}^{L^{2}} by

A~v\displaystyle\tilde{A}v^{*} :=(Av),v,\displaystyle:=\left(A^{*}v\right)^{*},\quad v\in\mathcal{H},
A~f\displaystyle\tilde{A}f :=0if fΓ1,𝐌L2\displaystyle:=0\qquad\text{if }f\in\mathcal{H}^{*\perp}\subset\Gamma_{1,{\bf M}}^{L^{2}}
Lemma 2.2.

Let AA be a bounded operator AA on \mathcal{H}. Define KA:𝕊(M)×𝕊(M)K_{A}:\mathbb{S}(M)\times\mathbb{S}(M)\to\mathbb{C} by

KA(v,u)=u|Av=v|A~u𝕊,u,v𝕊(M).K_{A}(v,u)=\left\langle u|Av\right\rangle=\left\langle v^{*}|\tilde{A}u^{*}\right\rangle_{\mathbb{S}},\quad u,v\in\mathbb{S}(M).

Then KAK_{A} is the integral kernel of A~\tilde{A}, i.e.,

(A~f)(v)=𝕊(𝐌)dμ𝕊(u)KA(v,u)f(u),fΓ1,𝐌L2,a.e. v𝕊(M).\left(\tilde{A}f\right)(v)=\int_{\mathbb{S}({\bf M})}{\rm d}\mu_{\mathbb{S}}(u)K_{A}(v,u)f(u),\quad\forall f\in\Gamma_{1,{\bf M}}^{L^{2}},\ \text{a.e. }v\in\mathbb{S}(M).

Especially, the kernel of 𝐩=|vv|𝐌\mathbf{p}=|v\rangle\langle v|\in{\bf M} is

K𝐩(v,v′′)=v′′|𝐩v=v′′|vv|v.K_{\mathbf{p}}(v^{\prime},v^{\prime\prime})=\left\langle v^{\prime\prime}|\mathbf{p}v^{\prime}\right\rangle=\left\langle v^{\prime\prime}|v\right\rangle\left\langle v|v^{\prime}\right\rangle. (2.4)
Proof.

Since A~f=A~Ef\tilde{A}f=\tilde{A}E_{\mathcal{H}^{*}}f, we can choose ww\in\mathcal{H} such that w=Efw^{*}=E_{\mathcal{H}^{*}}f.

𝕊(𝐌)dμ𝕊(u)u|Avw(u)\displaystyle\int_{\mathbb{S}({\bf M})}{\rm d}\mu_{\mathbb{S}}(u)\left\langle u|Av\right\rangle w^{*}(u)
=𝕊(𝐌)dμ𝕊(u)w|uu|Av\displaystyle\qquad=\int_{\mathbb{S}({\bf M})}{\rm d}\mu_{\mathbb{S}}(u)\left\langle w|u\right\rangle\left\langle u|Av\right\rangle
=w|Av=Aw|v=(Aw)(v)=(A~w)(v)=(A~f)(v)\displaystyle\qquad=\left\langle w|Av\right\rangle=\left\langle A^{*}w|v\right\rangle=\left(A^{*}w\right)^{*}(v)=\left(\tilde{A}w^{*}\right)(v)=\left(\tilde{A}f\right)(v)

Fix HL(𝐌)H\in L^{\infty}({\bf M}). For tt\in\mathbb{R}, define the bounded operator 𝒬t(F)\mathcal{Q}_{t}(F) and 𝒬~t(F)\tilde{\mathcal{Q}}_{t}(F) on \mathcal{H} and Γ1,𝐌L2\Gamma_{1,{\bf M}}^{L^{2}} respectively, by

𝒬t(F):=eit𝒬(H)𝒬(F)eit𝒬(H),𝒬~t(F):=eit𝒬~(H)𝒬~(F)eit𝒬~(H).\mathcal{Q}_{t}(F):=e^{\mathrm{i}t\mathcal{Q}(H)}\mathcal{Q}(F)e^{-\mathrm{i}t\mathcal{Q}(H)},\quad\tilde{\mathcal{Q}}_{t}(F):=e^{\mathrm{i}t\tilde{\mathcal{Q}}(H)}\tilde{\mathcal{Q}}(F)e^{-\mathrm{i}t\tilde{\mathcal{Q}}(H)}.

Note that

𝒬~t(F)=𝕊(𝐌)F~(v)𝐩𝐫t(v)dμ𝕊(v)=𝕊(𝐌)F~(v)𝐩𝐫(vt)dμ𝕊(v),\tilde{\mathcal{Q}}_{t}(F)=\int_{\mathbb{S}({\bf M})}\tilde{F}(v){\bf pr}_{t}(v^{*}){\rm d}\mu_{\mathbb{S}}(v)=\int_{\mathbb{S}({\bf M})}\tilde{F}(v){\bf pr}(v_{t}^{*}){\rm d}\mu_{\mathbb{S}}(v), (2.5)

where

𝐩𝐫t(v):=eit𝒬~(H)𝐩𝐫(v)eit𝒬~(H),vt:=eit𝒬(H)v,{\bf pr}_{t}(v^{*}):=e^{\mathrm{i}t\tilde{\mathcal{Q}}(H)}{\bf pr}(v^{*})e^{-\mathrm{i}t\tilde{\mathcal{Q}}(H)},\quad v_{t}:=e^{\mathrm{i}t\mathcal{Q}(H)}v,
𝐩t:=eit𝒬(H)𝐩eit𝒬(H).\mathbf{p}_{t}:=e^{\mathrm{i}t\mathcal{Q}(H)}\mathbf{p}e^{-\mathrm{i}t\mathcal{Q}(H)}.
Lemma 2.3.

For FL(𝐌,)F\in L^{\infty}({\bf M},\mathbb{C}),

(𝒬~t(F)s)(v)=𝕊(𝐌)Kt(v,u)s(u)dμ𝕊(u),sΓ1,𝐌L2,\left(\tilde{\mathcal{Q}}_{t}(F)s\right)(v)=\int_{\mathbb{S}({\bf M})}K_{t}(v,u)s(u){\rm d}\mu_{\mathbb{S}}(u),\qquad s\in\Gamma_{1,{\bf M}}^{L^{2}},

where

Kt(v,u)\displaystyle K_{t}(v,u) =v|𝒬~t(F)u=u|𝒬t(F)v\displaystyle=\left\langle v^{*}|\tilde{\mathcal{Q}}_{t}(F)u^{*}\right\rangle=\left\langle u|\mathcal{Q}_{t}(F)v\right\rangle
=𝕊(𝐌)dμ𝕊(x)F~(x)u|xtxt|v=𝐌dμ(𝐩)F(𝐩)u|𝐩tv\displaystyle=\int_{\mathbb{S}({\bf M})}{\rm d}\mu_{\mathbb{S}}(x)\tilde{F}(x)\left\langle u|x_{t}\right\rangle\left\langle x_{t}|v\right\rangle=\int_{{\bf M}}{\rm d}\mu(\mathbf{p})F(\mathbf{p})\left\langle u|\mathbf{p}_{t}v\right\rangle
Proof.

By (2.5),

(Q~t(F)s)(v)\displaystyle\left(\tilde{Q}_{t}(F)s\right)(v) =𝕊(𝐌)F~(u)(𝐩𝐫t(u)s)(v)dμ𝕊(u)\displaystyle=\int_{\mathbb{S}({\bf M})}\tilde{F}(u)\left({\bf pr}_{t}(u^{*})s\right)(v){\rm d}\mu_{\mathbb{S}}(u)
=𝕊(𝐌)dμ𝕊(u)F~(u)ut(v)ut|s𝕊\displaystyle=\int_{\mathbb{S}({\bf M})}{\rm d}\mu_{\mathbb{S}}(u)\tilde{F}(u)u_{t}^{*}(v)\left\langle u_{t}^{*}|s\right\rangle_{\mathbb{S}}
=𝕊(𝐌)dμ𝕊(u)F~(u)ut(v)𝕊(𝐌)dμ𝕊(x)ut(x)¯s(x)\displaystyle=\int_{\mathbb{S}({\bf M})}{\rm d}\mu_{\mathbb{S}}(u)\tilde{F}(u)u_{t}^{*}(v)\int_{\mathbb{S}({\bf M})}{\rm d}\mu_{\mathbb{S}}(x)\overline{u_{t}^{*}(x)}s(x)
=𝕊(𝐌)dμ𝕊(x)(𝕊(𝐌)dμ𝕊(u)F~(u)x|utut|v)s(x)\displaystyle=\int_{\mathbb{S}({\bf M})}{\rm d}\mu_{\mathbb{S}}(x)\left(\int_{\mathbb{S}({\bf M})}{\rm d}\mu_{\mathbb{S}}(u)\tilde{F}(u)\left\langle x|u_{t}\right\rangle\left\langle u_{t}|v\right\rangle\right)s(x)
=𝕊(𝐌)dμ𝕊(x)(𝕊(𝐌)dμ𝕊(u)F~(u)v|utut|x)s(x)\displaystyle=\int_{\mathbb{S}({\bf M})}{\rm d}\mu_{\mathbb{S}}(x)\left(\int_{\mathbb{S}({\bf M})}{\rm d}\mu_{\mathbb{S}}(u)\tilde{F}(u)\left\langle v^{*}|u_{t}^{*}\right\rangle\left\langle u_{t}^{*}|x^{*}\right\rangle\right)s(x)
=𝕊(𝐌)dμ𝕊(x)v|𝒬~t(F)xs(x).\displaystyle=\int_{\mathbb{S}({\bf M})}{\rm d}\mu_{\mathbb{S}}(x)\left\langle v^{*}|\tilde{\mathcal{Q}}_{t}(F)x^{*}\right\rangle s(x).

Lemma 2.4.

For t1,,tNt_{1},...,t_{N}\in\mathbb{R}, and H,F1,,FNL(𝐌,)H,F_{1},...,F_{N}\in L^{\infty}({\bf M},\mathbb{C}),

Tr𝒬~t1(F1)𝒬~tN(FN)=𝐌dμ(𝐩1)𝐌dμ(𝐩N)Tr(𝐩1,t1𝐩N,tN)j=1NFj(𝐩j),{\rm Tr}\tilde{\mathcal{Q}}_{t_{1}}(F_{1})\cdots\tilde{\mathcal{Q}}_{t_{N}}(F_{N})=\int_{{\bf M}}{\rm d}\mu(\mathbf{p}_{1})\cdots\int_{{\bf M}}{\rm d}\mu(\mathbf{p}_{N}){\rm Tr}\left(\mathbf{p}_{1,t_{1}}\cdots\mathbf{p}_{N,t_{N}}\right)\prod_{j=1}^{N}F_{j}(\mathbf{p}_{j}), (2.6)

where

𝐩j,tj:=eitj𝒬~(H)𝐩jeitj𝒬~(H).\mathbf{p}_{j,t_{j}}:=e^{\mathrm{i}t_{j}\tilde{\mathcal{Q}}(H)}\mathbf{p}_{j}e^{-\mathrm{i}t_{j}\tilde{\mathcal{Q}}(H)}.
Proof.

Let vN+1:=v1v_{N+1}:=v_{1}. Then by Lemma 2.3,

Tr𝒬~t1(F1)𝒬~tN(FN)\displaystyle{\rm Tr}\tilde{\mathcal{Q}}_{t_{1}}(F_{1})\cdots\tilde{\mathcal{Q}}_{t_{N}}(F_{N})
=𝕊(𝐌)dμ𝕊(v1)𝕊(𝐌)dμ𝕊(vN)j=1N𝐌dμ(𝐩j)Fj(𝐩j)vj|𝐩j,tjvj+1\displaystyle=\int_{\mathbb{S}({\bf M})}{\rm d}\mu_{\mathbb{S}}(v_{1})\cdots\int_{\mathbb{S}({\bf M})}{\rm d}\mu_{\mathbb{S}}(v_{N})\prod_{j=1}^{N}\int_{{\bf M}}{\rm d}\mu(\mathbf{p}_{j})\,F_{j}(\mathbf{p}_{j})\left\langle v_{j}|\mathbf{p}_{j,t_{j}}v_{j+1}\right\rangle
=𝐌dμ(𝐩1)𝐌dμ(𝐩N)Tr(𝐩1,t1𝐩N,tN)j=1NFj(𝐩j).\displaystyle=\int_{{\bf M}}{\rm d}\mu(\mathbf{p}_{1})\cdots\int_{{\bf M}}{\rm d}\mu(\mathbf{p}_{N}){\rm Tr}\left(\mathbf{p}_{1,t_{1}}\cdots\mathbf{p}_{N,t_{N}}\right)\prod_{j=1}^{N}F_{j}(\mathbf{p}_{j}).

3 Kähler manifold

In the following sections, we assume the

Assumption 3.1.

(1) 𝐌{\bf M} is a (finite-dimensional) Kähler submanifold of \mathbb{P}\mathcal{H},

(2) 𝐌{\bf M} is complete as a Riemannian manifold (or equivalently, complete as a metric space),

(3) For any fΓ1,𝐌L2f\in\Gamma_{1,{\bf M}}^{L^{2}}, ff\in\mathcal{H}^{*} if and only if ff is holomorphic.

(4) Let vol{\rm vol} be the volume form on 𝐌{\bf M} (as a Riemannian manifold). Then there exists a constant C>0C>0 s.t. the measure μ:=C|vol|\mu:=C\left|{\rm vol}\right| satisfies (𝐌,μ)Coh()({\bf M},\mu)\in{\rm Coh}(\mathcal{H}), i.e., 𝐌𝐩dμ(𝐩)=I\int_{{\bf M}}\mathbf{p}{\rm d}\mu(\mathbf{p})=I.

In the following we will explain the precise meaning of these assumptions.

Note that \mathbb{P}\mathcal{H} has the natural topology induced by that of \mathcal{H}, and so 𝐌{\bf M}\subset\mathbb{P}\mathcal{H} has the topology as a subspace of \mathbb{P}\mathcal{H}. Assume that there exist open sets UιnU_{\iota}\subset\mathbb{C}^{n} (ι\iota\in\mathcal{I}, some index set), and holomorphic maps ψι:Uι×:={0}\psi_{\iota}:U_{\iota}\to\mathcal{H}^{\times}:=\mathcal{H}\setminus\{0\} (ι\iota\in\mathcal{I}) such that

(1) {𝐩𝐫(ψι(Uι))}ι\left\{{\bf pr}\left(\psi_{\iota}(U_{\iota})\right)\right\}_{\iota\in\mathcal{I}} is an open cover of 𝐌{\bf M}.

(2) 𝐩𝐫ψι{\bf pr}\circ\psi_{\iota} is injective for all ι\iota\in\mathcal{I}.

Then we find that {𝐩𝐫ψι}ι\left\{{\bf pr}\circ\psi_{\iota}\right\}_{\iota\in\mathcal{I}} gives an atlas of 𝐌{\bf M} as a complex manifold. We assume without loss of generality that 0Uι0\in U_{\iota} and ψι(0)|ψι(z)=1\left\langle\psi_{\iota}(0)|\psi_{\iota}(z)\right\rangle=1 for all zUιz\in U_{\iota}, ι\iota\in\mathcal{I}. We write Uι:=𝐩𝐫ψι(Uι)𝐌U_{\iota}^{\prime}:={\bf pr}\circ\psi_{\iota}(U_{\iota})\subset{\bf M}, and locally identify UιU_{\iota}^{\prime} with UιU_{\iota}; We use the coordinates z=(z1,,zn)z=(z_{1},...,z_{n}) on UιU_{\iota} also as coordinates on UιU_{\iota}^{\prime}.

Let d{\rm d} denote the exterior derivative on 𝐌{\bf M}, which is decomposed to the holomorphic and antiholomorphic parts: d=d+d′′{\rm d}={{\rm d}^{\prime}}+{{\rm d}^{\prime\prime}}. For example, on UιU_{\iota}, for fC(Uι,)f\in C^{\infty}(U_{\iota},\mathbb{C}), df{{\rm d}^{\prime}}f and d′′f{{\rm d}^{\prime\prime}}f are explicitly written as

df=kzkfdzk,d′′f=kz¯kfdz¯k.{{\rm d}^{\prime}}f=\sum_{k}\frac{\partial}{\partial z_{k}}fdz_{k},\qquad{{\rm d}^{\prime\prime}}f=\sum_{k}\frac{\partial}{\partial\overline{z}_{k}}fd\overline{z}_{k}.

Define 𝐡ι:Uι{\bf h}_{\iota}:U_{\iota}\to\mathbb{R} by

𝐡ι(z):=ψι(z)2,zUι{\bf h}_{\iota}(z):=\left\|\psi_{\iota}(z)\right\|^{2},\qquad z\in U_{\iota}

Define the 2-form ω\omega on UιU_{\iota}by

ω:=id′′dlog𝐡ι.\omega:=-\mathrm{i}\,{\rm d}^{\prime\prime}{\rm d}^{\prime}\log{\bf h}_{\iota}.

It turns out that ω\omega becomes a globally defined closed 2-form on 𝐌{\bf M}, and hence (𝐌,ω)({\bf M},\omega) is a symplectic manifold. In fact, ω\omega is naturally defined on whole projective space \mathbb{P}\mathcal{H} as follows: Define 𝐡:×{\bf h}:\mathcal{H}^{\times}\to\mathbb{R} by 𝐡(v):=v2{\bf h}(v):=\left\|v\right\|^{2}. Then the 2-form ω:=id′′dlog𝐡\omega:=-\mathrm{i}\,{\rm d}^{\prime\prime}{\rm d}^{\prime}\log{\bf h} on ×\mathcal{H}^{\times} is defined even when dim=\dim\mathcal{H}=\infty. Since ω\omega is invariant under the action of ×:={0}\mathbb{C}^{\times}:=\mathbb{C}\setminus\{0\} on ×\mathcal{H}^{\times}, ω\omega can be viewed as a 2-form on ×/×\mathbb{P}\mathcal{H}\cong\mathcal{H}^{\times}/\mathbb{C}^{\times}.

For tangent vector fields X,YX,Y on 𝐌{\bf M}, let

g(X,Y):=ω(X,JY),g(X,Y):=\omega(X,JY),

where JJ is the complex structure on 𝐌{\bf M}. Then gg becomes a Riemannian metric on 𝐌{\bf M}, and moreover we find that (𝐌,g,ω)({\bf M},g,\omega) is a Kähler manifold. We call 𝐡ι{\bf h}_{\iota} a Kähler potential on UιU_{\iota}^{\prime}.

Therefore, we find that if 𝐌{\bf M} is a (finite-dimensional) complex submanifold of \mathbb{P}\mathcal{H}, 𝐌{\bf M} satisfies Assumption 3.1 (1), even when dim=\dim\mathcal{H}=\infty.

Assumption 3.1 (2) says that every geodesic line of 𝐌{\bf M} can be extended for arbitrarily large values of its canonical parameter. This is equivalent to say that 𝐌{\bf M} is a complete metric space with respect to the distance function dd induced by the Riemannian metric gg.

Assumption 3.1 (3) says that for any fΓ1,𝐌L2f\in\Gamma_{1,{\bf M}}^{L^{2}} if fψι:Uιf\circ\psi_{\iota}:U_{\iota}\to\mathbb{C} is holomorphic for all ι\iota\in\mathcal{I}, then ff\in\mathcal{H}^{*}. (The converse always holds.)

3.1 Line bundle

Let fΓ,𝐌f\in\Gamma_{\ell,{\bf M}} (\ell\in\mathbb{Z}) and v𝐩𝐫1(𝐌)v\in{\bf pr}^{-1}({\bf M}). We define the value of ff at 𝐩=𝐩𝐫(v)𝐌\mathbf{p}={\bf pr}(v)\in{\bf M}, denoted f(𝐩)=f(𝐩𝐫(v))f(\mathbf{p})=f({\bf pr}(v)), to be

f(𝐩):=f|ran(𝐩){0},equivalently,f(𝐩𝐫(v)):=f|×v,f(\mathbf{p}):=f|_{\mathrm{ran}(\mathbf{p})\setminus\{0\}},\quad\text{equivalently,}\quad f({\bf pr}(v)):=f|_{\mathbb{C}^{\times}v},

that is, f(𝐩𝐫(v))f({\bf pr}(v)) is the function ×v\mathbb{C}^{\times}v\to\mathbb{C} defined by

f(𝐩𝐫(v))(αv):=f(αv)=αf(v),α×.f({\bf pr}(v))(\alpha v):=f(\alpha v)=\alpha^{\ell}f(v),\qquad\alpha\in\mathbb{C}^{\times}.

Let

𝒪,𝐌:={f(𝐩)|fΓ,𝐌,𝐩𝐌},\mathscr{O}_{\ell,{\bf M}}:=\left\{f(\mathbf{p})|f\in\Gamma_{\ell,{\bf M}},\ \mathbf{p}\in{\bf M}\right\},

then the natural projection 𝒪,𝐌𝐌\mathscr{O}_{\ell,{\bf M}}\to{\bf M}, f(𝐩)𝐩f(\mathbf{p})\mapsto\mathbf{p} defines a complex line bundle over 𝐌{\bf M}, where each fΓ,𝐌f\in\Gamma_{\ell,{\bf M}} is a section of the line bundle. The space 𝐀r(𝒪,𝐌){\bf A}^{r}(\mathscr{O}_{\ell,{\bf M}}) of 𝒪,𝐌\mathscr{O}_{\ell,{\bf M}}-valued rr-forms are usually defined. For fΓ,𝐌f\in\Gamma_{\ell,{\bf M}}, let supp𝐌(f){\rm supp}_{{\bf M}}(f) denote the support of ff as a map 𝐌𝒪,𝐌{\bf M}\to\mathscr{O}_{\ell,{\bf M}}, and then for any α𝐀r(𝒪,𝐌)\alpha\in{\bf A}^{r}(\mathscr{O}_{\ell,{\bf M}}), the support supp𝐌(α)𝐌{\rm supp}_{{\bf M}}(\alpha)\subset{\bf M} of α\alpha is naturally defined.

Here, recall Lemma 2.2; The integral kernel KAK_{A} of an operator A~\tilde{A} on Γ1,𝐌L2\Gamma_{1,{\bf M}}^{L^{2}} was a \mathbb{C}-valued function on 𝕊(𝐌)×𝕊(𝐌)\mathbb{S}({\bf M})\times\mathbb{S}({\bf M}) there. However, if Γ1,𝐌L2\Gamma_{1,{\bf M}}^{L^{2}} is viewed as a space of sections s:𝐌𝒪1,𝐌s:{\bf M}\to\mathscr{O}_{1,{\bf M}}, the an integral kernel KK of an operator on Γ1,𝐌L2\Gamma_{1,{\bf M}}^{L^{2}} should be a map such that K(𝐩1,𝐩2)Hom(𝒪1,𝐌,𝐩2,𝒪1,𝐌,𝐩1)K(\mathbf{p}_{1},\mathbf{p}_{2})\in\mathrm{Hom}(\mathscr{O}_{1,{\bf M},\mathbf{p}_{2}},\mathscr{O}_{1,{\bf M},\mathbf{p}_{1}}) for all 𝐩1,𝐩2𝐌\mathbf{p}_{1},\mathbf{p}_{2}\in{\bf M}, where 𝒪1,𝐌,𝐩\mathscr{O}_{1,{\bf M},\mathbf{p}} is the fiber of the line bundle 𝒪1,𝐌\mathscr{O}_{1,{\bf M}} at 𝐩𝐌\mathbf{p}\in{\bf M}; Equivalently, KK is a section of the external tensor product bundle 𝒪1,𝐌𝒪1,𝐌𝐌×𝐌\mathscr{O}_{1,{\bf M}}\boxtimes\mathscr{O}_{1,{\bf M}}^{*}\to{\bf M}\times{\bf M}. Note that 𝒪1,𝐌,𝐩\mathscr{O}_{1,{\bf M},\mathbf{p}} is naturally identified with the dual space of ran(𝐩)\mathrm{ran}(\mathbf{p}). Hence we can define KA(𝐩1,𝐩2)Hom(𝒪1,𝐌,𝐩2,𝒪1,𝐌,𝐩1)K_{A}(\mathbf{p}_{1},\mathbf{p}_{2})\in\mathrm{Hom}(\mathscr{O}_{1,{\bf M},\mathbf{p}_{2}},\mathscr{O}_{1,{\bf M},\mathbf{p}_{1}}) as follows. For any v2𝒪1,𝐌,𝐩2=ran(𝐩2)v_{2}^{*}\in\mathscr{O}_{1,{\bf M},\mathbf{p}_{2}}=\mathrm{ran}(\mathbf{p}_{2})^{*} with v2ran(𝐩2)v_{2}\in\mathrm{ran}(\mathbf{p}_{2}), define KA(𝐩1,𝐩2)v2𝒪1,𝐌,𝐩1=ran(𝐩1)K_{A}(\mathbf{p}_{1},\mathbf{p}_{2})v_{2}^{*}\in\mathscr{O}_{1,{\bf M},\mathbf{p}_{1}}=\mathrm{ran}(\mathbf{p}_{1})^{*} by

(KA(𝐩1,𝐩2)v2)(v1):=v2|Av1=KA(v1,v2),v1ran(𝐩1).\left(K_{A}(\mathbf{p}_{1},\mathbf{p}_{2})v_{2}^{*}\right)(v_{1}):=\left\langle v_{2}|Av_{1}\right\rangle=K_{A}(v_{1},v_{2}),\quad v_{1}\in\mathrm{ran}(\mathbf{p}_{1}). (3.1)

Sometimes we simply write A(𝐩1,𝐩2)A(\mathbf{p}_{1},\mathbf{p}_{2}) (resp. A(v1,v2)A(v_{1},v_{2})) for KA(𝐩1,𝐩2)K_{A}(\mathbf{p}_{1},\mathbf{p}_{2}) (resp. KA(v1,v2)K_{A}(v_{1},v_{2})). Especially we have

(E(𝐩1,𝐩2)v2)(v1)=v2|v1=E(v1,v2),v1ran(𝐩1).\left(E_{\mathcal{H}^{*}}(\mathbf{p}_{1},\mathbf{p}_{2})v_{2}^{*}\right)(v_{1})=\left\langle v_{2}|v_{1}\right\rangle=E_{\mathcal{H}^{*}}(v_{1},v_{2}),\quad v_{1}\in\mathrm{ran}(\mathbf{p}_{1}). (3.2)

Let f1,f2Γ,𝐌f_{1},f_{2}\in\Gamma_{\ell,{\bf M}} (\ell\in\mathbb{Z}). Define f1|f2pt:𝐌\left\langle f_{1}|f_{2}\right\rangle_{{\rm pt}}:{\bf M}\to\mathbb{C} by

f1|f2pt(𝐩𝐫(v)):=f1(vv)¯f2(vv)=v2f1(v)¯f2(v),v𝐩𝐫1(𝐌)\left\langle f_{1}|f_{2}\right\rangle_{{\rm pt}}({\bf pr}(v)):=\overline{f_{1}\left(\frac{v}{\left\|v\right\|}\right)}f_{2}\left(\frac{v}{\left\|v\right\|}\right)=\left\|v\right\|^{-2\ell}\overline{f_{1}\left(v\right)}f_{2}\left(v\right),\quad v\in{\bf pr}^{-1}({\bf M})

so that if f1,f2Γ,𝐌L2f_{1},f_{2}\in\Gamma_{\ell,{\bf M}}^{L^{2}} then

f1|f2𝕊=𝐌f1|f2ptCvol.\left\langle f_{1}|f_{2}\right\rangle_{\mathbb{S}}=\int_{{\bf M}}\left\langle f_{1}|f_{2}\right\rangle_{{\rm pt}}C{\rm vol}.

|pt\left\langle\cdot|\cdot\right\rangle_{{\rm pt}} is called the (pointwise) Hermitian metric of the line bundle 𝒪,𝐌\mathscr{O}_{\ell,{\bf M}}.

Let eι:=ψι(0)e_{\iota}:=\psi_{\iota}(0), then eιe_{\iota}^{*} is a local holomorphic section of 𝒪1,𝐌\mathscr{O}_{1,{\bf M}} on UιU_{\iota}; Precisely, if we identify the fiber of 𝒪1,𝐌\mathscr{O}_{1,{\bf M}} over 𝐩𝐌\mathbf{p}\in{\bf M} with ran(𝐩)\mathrm{ran}(\mathbf{p})^{*}, the value eι(𝐩)e_{\iota}^{*}(\mathbf{p}) of the section eιΓ1,𝐌e_{\iota}^{*}\in\Gamma_{1,{\bf M}} at 𝐩Uι\mathbf{p}\in U_{\iota} is the function ran(𝐩)\mathrm{ran}(\mathbf{p})\to\mathbb{C}, veι|vv\mapsto\left\langle e_{\iota}|v\right\rangle.

Define hι:Uιh_{\iota}:U_{\iota}^{\prime}\to\mathbb{R} by

hι(z):=eι|eιpt(z),zUι.h_{\iota}(z):=\left\langle e_{\iota}^{*}|e_{\iota}^{*}\right\rangle_{{\rm pt}}(z),\qquad z\in U_{\iota}^{\prime}.

Without loss of generality, assume ψι(0)|ψι(z)=1\left\langle\psi_{\iota}(0)|\psi_{\iota}(z)\right\rangle=1, and we find

hι(z)=ψι(z)2eι(ψι(z))¯eι(ψι(z))=ψι(z)2eι|ψι(z)¯eι|ψι(z)=ψι(z)2=1𝐡ι(z).h_{\iota}(z)=\left\|\psi_{\iota}(z)\right\|^{-2}\overline{e_{\iota}^{*}\left(\psi_{\iota}(z)\right)}e_{\iota}^{*}\left(\psi_{\iota}(z)\right)=\left\|\psi_{\iota}(z)\right\|^{-2}\overline{\left\langle e_{\iota}|\psi_{\iota}(z)\right\rangle}\left\langle e_{\iota}|\psi_{\iota}(z)\right\rangle=\left\|\psi_{\iota}(z)\right\|^{-2}=\frac{1}{{\bf h}_{\iota}(z)}.

Let Γ,𝐌\Gamma_{\ell,{\bf M}}^{\infty} (\ell\in\mathbb{Z}) denote the subspace of Γ,𝐌\Gamma_{\ell,{\bf M}} which consists of the smooth functions. Any section sΓ1,𝐌s\in\Gamma_{1,{\bf M}}^{\infty} is uniquely expressed locally on Uι𝐌U_{\iota}^{\prime}\subset{\bf M} by

s(v)=f(v)eι(v)=f(v)eι|v,fΓ0,𝐌,v×ψι(Uι),s(v)=f(v)e_{\iota}^{*}(v)=f(v)\left\langle e_{\iota}|v\right\rangle,\qquad f\in\Gamma_{0,{\bf M}}^{\infty},\ v\in\mathbb{C}^{\times}\psi_{\iota}(U_{\iota}),

which is simply written as s=feιs=fe_{\iota}^{*}. The antiholomorphic exterior derivative d′′s{{\rm d}^{\prime\prime}}s of s=feιs=fe_{\iota}^{*} is defined on UιU_{\iota}^{\prime} by

d′′(feι):=(d′′f)eι.{{\rm d}^{\prime\prime}}\left(fe_{\iota}^{*}\right):=\left({{\rm d}^{\prime\prime}}f\right)\otimes e_{\iota}^{*}.

We find that d′′s𝐀1(𝒪1,𝐌){{\rm d}^{\prime\prime}}s\in{\bf A}^{1}(\mathscr{O}_{1,{\bf M}}) is globally well-defined on 𝐌{\bf M}. This is naturally extended to any 𝒪1,𝐌\mathscr{O}_{1,{\bf M}}-valued rr-forms d′′:𝐀r(𝒪1,𝐌)𝐀r+1(𝒪1,𝐌){{\rm d}^{\prime\prime}}:{\bf A}^{r}(\mathscr{O}_{1,{\bf M}})\to{\bf A}^{r+1}(\mathscr{O}_{1,{\bf M}}); More precisely, let 𝐀(p,q)(𝒪1,𝐌){\bf A}^{(p,q)}(\mathscr{O}_{1,{\bf M}}) denote the space of 𝒪1,𝐌\mathscr{O}_{1,{\bf M}}-valued differential forms of type (p,q)(p,q), then d′′:𝐀(p,q)(𝒪1,𝐌)𝐀(p,q+1)(𝒪1,𝐌){{\rm d}^{\prime\prime}}:{\bf A}^{(p,q)}(\mathscr{O}_{1,{\bf M}})\to{\bf A}^{(p,q+1)}(\mathscr{O}_{1,{\bf M}}).

The local Chern connection form θ\theta of the line bundle 𝒪1,𝐌\mathscr{O}_{1,{\bf M}} on UιU_{\iota}\subset\mathbb{C} is defined by

θ:=dloghι=dlog𝐡ι=k(hι1zkhι)dzk.\theta:={\rm d}^{\prime}\log h_{\iota}=-{\rm d}^{\prime}\log{\bf h}_{\iota}=\sum_{k}\left(h_{\iota}^{-1}\frac{\partial}{\partial z_{k}}h_{\iota}\right){\rm d}z_{k}.

The Chern connection :𝐀0(𝒪1,𝐌)𝐀1(𝒪1,𝐌)\nabla:{\bf A}^{0}(\mathscr{O}_{1,{\bf M}})\to{\bf A}^{1}(\mathscr{O}_{1,{\bf M}}) is defined as follows: for any section sΓ1,𝐌s\in\Gamma_{1,{\bf M}}^{\infty} which is s=feιs=fe_{\iota}^{*} on Uι𝐌U_{\iota}^{\prime}\subset{\bf M}, s\nabla s is defined locally on UιU_{\iota}^{\prime} as

s:=(df)eι+feι,eι:=θeι.\nabla s:=\left({\rm d}f\right)\otimes e_{\iota}^{*}+f\nabla e_{\iota}^{*},\qquad\nabla e_{\iota}^{*}:=\theta\otimes e_{\iota}^{*}.

The Chern connection \nabla is Hermitian, i.e., for any vector field XX,

Xs1|s2pt=Xs1|s2pt+s1|Xs2pt,s1,s2Γ1,𝐌.X\left\langle s_{1}|s_{2}\right\rangle_{{\rm pt}}=\left\langle\nabla_{X}s_{1}|s_{2}\right\rangle_{{\rm pt}}+\left\langle s_{1}|\nabla_{X}s_{2}\right\rangle_{{\rm pt}},\qquad s_{1},s_{2}\in\Gamma_{1,{\bf M}}^{\infty}.

Hence, for any piecewise smooth curve cc on 𝐌{\bf M}, the parallel transport //(c)//(c) along cc is unitary w.r.t. the Hermitian metric |pt\left\langle\cdot|\cdot\right\rangle_{{\rm pt}}. Especially, the holonomy group at any point 𝐩𝐌\mathbf{p}\in{\bf M} w.r.t. \nabla is U(1)U(1). Let ψ(z)¯:=ψ(z)/ψ(z)\underline{\psi(z)}:=\psi(z)/\left\|\psi(z)\right\|. The normalized frame eι¯Γ1,𝐌\underline{e_{\iota}^{*}}\in\Gamma_{1,{\bf M}} on UιU_{\iota} is determined by

eι¯(ψ(z)¯)=1=h1/2(z)eι(ψ(z)¯),z,\underline{e_{\iota}^{*}}(\underline{\psi(z)})=1=h^{-1/2}(z)e_{\iota}^{*}(\underline{\psi(z)}),\qquad z\in\mathbb{C},

which is written eι¯:=hι1/2eι\underline{e_{\iota}^{*}}:=h_{\iota}^{-1/2}e_{\iota}^{*} in short. (Here, recall eι(ψι(z))=ψι(0)|ψι(z)=1e_{\iota}^{*}(\psi_{\iota}(z))=\left\langle\psi_{\iota}(0)|\psi_{\iota}(z)\right\rangle=1 and hι(z)1=ψι(z)2=𝐡ι(z).h_{\iota}(z)^{-1}=\left\|\psi_{\iota}(z)\right\|^{2}={\bf h}_{\iota}(z).) Precisely, the value eι¯(𝐩)\underline{e_{\iota}^{*}}(\mathbf{p}) of the section eι¯\underline{e_{\iota}^{*}} at 𝐩Uι\mathbf{p}\in U_{\iota} is the function

ran(𝐩),ζψ(z)¯hι1/2(z)eι|ζψ(z)¯=ζ,zd,ψ(z)ran(𝐩),ζ.\mathrm{ran}(\mathbf{p})\to\mathbb{C},\qquad\zeta\underline{\psi(z)}\mapsto h_{\iota}^{-1/2}(z)\bigl{\langle}e_{\iota}^{*}|\zeta\underline{\psi(z)}\bigr{\rangle}=\zeta,\qquad z\in\mathbb{C}^{d},\ \psi(z)\in\mathrm{ran}(\mathbf{p}),\ \zeta\in\mathbb{C}.

For s=feι¯s=f\underline{e_{\iota}}^{*}, fC(Uι,)f\in C^{\infty}(U_{\iota},\mathbb{C}), we have

s=(df)eι¯+fiθnoreι¯,\nabla s=\left({\rm d}f\right)\otimes\underline{e_{\iota}}^{*}+f\mathrm{i}\theta_{{\rm nor}}\otimes\underline{e_{\iota}}^{*},

where

θnor:=12k[(xkloghι)dyk(ykloghι)dxk],zk=xk+iyk,\theta_{{\rm nor}}:=\frac{1}{2}\sum_{k}\left[\left(\frac{\partial}{\partial x_{k}}\log h_{\iota}\right){\rm d}y_{k}-\left(\frac{\partial}{\partial y_{k}}\log h_{\iota}\right){\rm d}x_{k}\right],\quad z_{k}=x_{k}+\mathrm{i}y_{k}, (3.3)

which is a \mathbb{R}-valued 1-form; Since i=𝔲(1)\mathrm{i}\mathbb{R}=\mathfrak{u}(1) (the Lie algebra of U(1)U(1)), this says that the Chern connection \nabla determines a connection on the associated U(1)U(1) principal bundle, explicitly written by iθnor\mathrm{i}\theta_{{\rm nor}}.

For s=feιs=fe_{\iota}^{*}, let

s:=(df)eι+f(θeι),\nabla^{\prime}s:=\left({\rm d}^{\prime}f\right)\otimes e_{\iota}^{*}+f\left(\theta\otimes e_{\iota}^{*}\right),

then we have

=+d′′.\nabla=\nabla^{\prime}+{{\rm d}^{\prime\prime}}.

We find that :𝐀0(𝒪1,𝐌)𝐀1(𝒪1,𝐌)\nabla:{\bf A}^{0}(\mathscr{O}_{1,{\bf M}})\to{\bf A}^{1}(\mathscr{O}_{1,{\bf M}}) and :𝐀(0,0)(𝒪1,𝐌)𝐀(1,0)(𝒪1,𝐌)\nabla^{\prime}:{\bf A}^{(0,0)}(\mathscr{O}_{1,{\bf M}})\to{\bf A}^{(1,0)}(\mathscr{O}_{1,{\bf M}}) are globally well-defined on 𝐌{\bf M}.

For general differential forms α𝐀r(𝒪1,𝐌)\alpha\in{\bf A}^{r}(\mathscr{O}_{1,{\bf M}}), α𝐀r+1(𝒪1,𝐌)\nabla\alpha\in{\bf A}^{r+1}(\mathscr{O}_{1,{\bf M}}) is defined by

α:=j(dαjsj+(1)rαjsj)\nabla\alpha:=\sum_{j}({\rm d}\alpha_{j}\otimes s_{j}+(-1)^{r}\alpha_{j}\wedge\nabla s_{j})

where α=jαjsj\alpha={\displaystyle\sum_{j}\alpha_{j}\otimes s_{j}} (αj𝐀r(𝐌,)\alpha_{j}\in{\bf A}^{r}({\bf M},\mathbb{C}), sjΓ1,𝐌s_{j}\in\Gamma_{1,{\bf M}}^{\infty}).

It is shown that there exists Θ𝐀2(𝐌,)\Theta\in{\bf A}^{2}({\bf M},\mathbb{C}) such that

2α=Θα,α𝐀k(𝒪1,𝐌),k=0,,n.\nabla^{2}\alpha=\Theta\wedge\alpha,\qquad\forall\alpha\in{\bf A}^{k}(\mathscr{O}_{1,{\bf M}}),\ k=0,...,n. (3.4)

Θ\Theta is called the curvature form of the Chern connection \nabla. Generally, the curvature form Θ\Theta is locally expressed by the Hermitian metric hh by

Θ=d′′dloghι,\Theta={\rm d}^{\prime\prime}{\rm d}^{\prime}\log h_{\iota},

and so we have

ω=iΘ\omega=\mathrm{i}\Theta (3.5)

in our case.

3.2 The Bochner-Kodaira-Nakano identity

Let Γ1,𝐌,c\Gamma_{1,{\bf M},{\rm c}}^{\infty} denote the space of compactly supported smooth sections of 𝒪1,𝐌\mathscr{O}_{1,{\bf M}}, and 𝐀cr(𝒪1,𝐌){\bf A}_{{\rm c}}^{r}(\mathscr{O}_{1,{\bf M}}) the space of compactly supported 𝒪1,𝐌\mathscr{O}_{1,{\bf M}}-valued rr-forms on 𝐌{\bf M}. The inner product |𝕊\left\langle\cdot|\cdot\right\rangle_{\mathbb{S}} is naturally extended to 𝐀cr(𝒪1,𝐌){\bf A}_{{\rm c}}^{r}(\mathscr{O}_{1,{\bf M}}), and the formal adjoint operators are usually defined on these spaces: \nabla^{*}, \nabla^{\prime*} and d′′{{\rm d}^{\prime\prime}}^{*}are the formal adjoints of \nabla, \nabla^{\prime} and d′′{{\rm d}^{\prime\prime}}, respectively. We use the notation ΔX:=XX+XX\Delta_{X}:=X^{*}X+XX^{*} for any operator XX on 𝐀cr(𝒪1,𝐌){\bf A}_{c}^{r}(\mathscr{O}_{1,{\bf M}}), if the formal adjoint XX^{*} of XX exists. For example,

Δd′′:=d′′d′′+d′′d′′,Δ:=+.\Delta_{{{\rm d}^{\prime\prime}}}:={{\rm d}^{\prime\prime}}^{*}{{\rm d}^{\prime\prime}}+{{\rm d}^{\prime\prime}}{{\rm d}^{\prime\prime}}^{*},\qquad{\Delta_{\text{\tiny$\nabla$}}}:=\nabla^{*}\nabla+\nabla\nabla^{*}.

The Lefschetz map L:𝐀(p,q)(𝒪1,𝐌)𝐀(p+1,q+1)(𝒪1,𝐌)L:{\bf A}^{(p,q)}(\mathscr{O}_{1,{\bf M}})\to{\bf A}^{(p+1,q+1)}(\mathscr{O}_{1,{\bf M}}) is defined by

L(α):=ωα.L(\alpha):=\omega\wedge\alpha.

Since ω=iΘ\omega=\mathrm{i}\Theta holds in our case, we have L(α)=iΘα=i2αL(\alpha)=\mathrm{i}\Theta\wedge\alpha=\mathrm{i}\nabla^{2}\alpha, i.e.,

L=i2.L=\mathrm{i}\nabla^{2}. (3.6)

Let PrP_{r} denote the natural projection onto 𝐀r(𝒪1,𝐌){\bf A}^{r}(\mathscr{O}_{1,{\bf M}}): If α=k=0dim𝐌αk\alpha=\sum_{k=0}^{\dim{\bf M}}\alpha_{k}, αk𝐀k(𝒪1,𝐌)\alpha_{k}\in{\bf A}^{k}(\mathscr{O}_{1,{\bf M}}), then Pr(α):=αrP_{r}(\alpha):=\alpha_{r}.

Lemma 3.2.

(See e.g. [1, Ch.5]; [7, Eq.(3.2.37)]) [L,L]=r=0n(nr)Pr[L^{*},L]=\sum_{r=0}^{n}(n-r)P_{r}, where n=dim𝐌n=\dim_{\mathbb{R}}{\bf M}.

Theorem 3.3.

(Bochner–Kodaira–Nakano identity) (See e.g. [1, Ch.5])

2Δd′′=Δ+[i2,L]2\Delta_{{\rm d}^{\prime\prime}}={\Delta_{\text{\tiny$\nabla$}}}+[\mathrm{i}\nabla^{2},L^{*}] (3.7)

By (3.6): L=i2L=\mathrm{i}\nabla^{2}, Lemma 3.2 and the Bochner–Kodaira–Nakano identity (3.7), we have

Corollary 3.4.

On the line bundle 𝒪1,𝐌\mathscr{O}_{1,{\bf M}} over 𝐌{\bf M}, we have

2Δd′′=Δ+[L,L]=Δr=0n(nr)Pr,n=dim𝐌.2\Delta_{{\rm d}^{\prime\prime}}={\Delta_{\text{\tiny$\nabla$}}}+[L,L^{*}]={\Delta_{\text{\tiny$\nabla$}}}-\sum_{r=0}^{n}(n-r)P_{r},\qquad n=\dim_{\mathbb{R}}{\bf M}.

Especially,

2Δd′′s=(Δn)s=(n)s,s𝐀0(𝒪1,𝐌)=Γ1,𝐌.2\Delta_{{\rm d}^{\prime\prime}}s=\left({\Delta_{\text{\tiny$\nabla$}}}-n\right)s=\left(\nabla^{*}\nabla-n\right)s,\qquad\forall s\in{\bf A}^{0}(\mathscr{O}_{1,{\bf M}})=\Gamma_{1,{\bf M}}^{\infty}.

Generally, the operators Δ{\Delta_{\text{\tiny$\nabla$}}} and Δd′′\Delta_{{\rm d}^{\prime\prime}} are shown to be essentially self-adjoint, and we write the self-adjoint extensions these operators again by Δ{\Delta_{\text{\tiny$\nabla$}}} and Δd′′\Delta_{{\rm d}^{\prime\prime}}, respectively.

Corollary 3.5.
E=s-lim𝝂exp[𝝂(n)],E_{\mathcal{H}^{*}}=\mathop{\mbox{\rm s-lim}}_{\boldsymbol{\nu}\to\infty}\exp\left[-\boldsymbol{\nu}\left(\nabla^{*}\nabla-n\right)\right],

where s-lim\mathop{\mbox{\rm s-lim}} denotes the limit in the strong topology on B(Γ1,𝐌L2)B(\Gamma_{1,{\bf M}}^{L^{2}}), the space of bounded operators on Γ1,𝐌L2\Gamma_{1,{\bf M}}^{L^{2}}.

We say that Δd′′\Delta_{{\rm d}^{\prime\prime}} has a spectral gap if inf(spec(Δd′′){0})>0\inf\left({\rm spec}(\Delta_{{\rm d}^{\prime\prime}})\setminus\{0\}\right)>0, where spec(Δd′′){\rm spec}(\Delta_{{\rm d}^{\prime\prime}}) is the spectrum of Δd′′\Delta_{{\rm d}^{\prime\prime}}; In other words, Δd′′α(IE)\Delta_{{\rm d}^{\prime\prime}}\geq\alpha(I-E_{\mathcal{H}^{*}}) for some α>0\alpha>0.

Corollary 3.6.

If Δd′′\Delta_{{\rm d}^{\prime\prime}} has a spectral gap, then

E=lim𝝂exp[𝝂(n)],E_{\mathcal{H}^{*}}=\lim_{\boldsymbol{\nu}\to\infty}\exp\left[-\boldsymbol{\nu}\left(\nabla^{*}\nabla-n\right)\right],

where lim\lim denotes the limit in operator norm on B(Γ1,𝐌L2)B(\Gamma_{1,{\bf M}}^{L^{2}}).

4 Asymptotic representation

Let \mathcal{L} be an arbitrary complex Hilbert space, and 𝒦\mathcal{K} be a closed subspace of \mathcal{L}. Let HH be a bounded self-adjoint operator on \mathcal{L}, and AA a possibly unbounded positive semidefinite operator on \mathcal{L} such that kerA=𝒦\ker A=\mathcal{K}\subset\mathcal{L}. Assume that AA has a spectral gap, i.e., there exists α>0\alpha>0 such that the spectrum of AA satisfies spec(A)(0,α)={\rm spec}(A)\cap(0,\alpha)=\emptyset, equivalently A2αAA^{2}\geq\alpha A, or AαE𝒦A\geq\alpha E_{\mathcal{K}}^{\perp} where E𝒦:=1E𝒦E_{\mathcal{K}}^{\perp}:=1-E_{\mathcal{K}}.

This section we show the following theorem:

Theorem 4.1.

For all t0t\geq 0,

lim𝝂et(𝝂A+iH)v\displaystyle\lim_{\boldsymbol{\nu}\to\infty}e^{-t\left(\boldsymbol{\nu}A+\mathrm{i}H\right)}v =eitE𝒦HE𝒦v,v𝒦=kerA,\displaystyle=e^{-\mathrm{i}tE_{\mathcal{K}}HE_{\mathcal{K}}}v,\qquad\forall v\in\mathcal{K}=\ker A,
lim𝝂et(𝝂A+iH)v\displaystyle\lim_{\boldsymbol{\nu}\to\infty}e^{-t\left(\boldsymbol{\nu}A+\mathrm{i}H\right)}v =0,v𝒦.\displaystyle=0,\qquad\forall v\in\mathcal{K}^{\perp}.

That is,

s-lim𝝂et(𝝂A+iH)=eitE𝒦HE𝒦E𝒦=E𝒦eitE𝒦HE𝒦E𝒦.\mathop{\mbox{\rm s-lim}}_{\boldsymbol{\nu}\to\infty}e^{-t\left(\boldsymbol{\nu}A+\mathrm{i}H\right)}=e^{-\mathrm{i}tE_{\mathcal{K}}HE_{\mathcal{K}}}E_{\mathcal{K}}=E_{\mathcal{K}}e^{-\mathrm{i}tE_{\mathcal{K}}HE_{\mathcal{K}}}E_{\mathcal{K}}.

Note that we can give a precise meaning to the operator et(𝝂A+iH)e^{-t\left(\boldsymbol{\nu}A+\mathrm{i}H\right)} as follows. We find that T𝝂:=𝝂A+iHT_{\boldsymbol{\nu}}:=\boldsymbol{\nu}A+\mathrm{i}H is a closed operator satisfying v|T𝝂v0\Re\langle v|T_{\boldsymbol{\nu}}v\rangle\geq 0 for all vdom(T𝝂)=dom(A)v\in\mathrm{dom}(T_{\boldsymbol{\nu}})=\mathrm{dom}(A). Hence T𝝂T_{\boldsymbol{\nu}} generates the strongly continuous contraction semigroup {etT𝝂|t0}\{e^{-tT_{\boldsymbol{\nu}}}|t\geq 0\} by the Hille–Yosida Theorem [9].

Recall the notations and assumptions in the previous section. Set

:=Γ1,𝐌L2,𝒦:=Γ1,𝐌L2,A:=2Δd′′Γ1,𝐌L2=n,n:=dim𝐌,\mathcal{L}:=\Gamma_{1,{\bf M}}^{L^{2}},\quad\mathcal{K}:=\mathcal{H}^{*}\subset\Gamma_{1,{\bf M}}^{L^{2}},\quad A:=2\Delta_{{\rm d}^{\prime\prime}}\upharpoonright_{\Gamma_{1,{\bf M}}^{L^{2}}}=\nabla^{*}\nabla-n,\quad n:=\dim_{\mathbb{R}}{\bf M},

in Theorem 4.1, then we have the following:

Corollary 4.2.

Assume that Δd′′\Delta_{{\rm d}^{\prime\prime}} has a spectral gap. Then for any “classical Hamiltonian” HL(𝐌,)H\in L^{\infty}({\bf M},\mathbb{R}), the “quantum time evolution” eit𝒬~(H)e^{-\mathrm{i}t\tilde{\mathcal{Q}}(H)} is expressed as

eit𝒬~(H)E=Eeit𝒬~(H)E=s-lim𝝂exp[t(𝝂(n)+iH)],t0,e^{-\mathrm{i}t\tilde{\mathcal{Q}}(H)}E_{\mathcal{H}^{*}}=E_{\mathcal{H}^{*}}e^{-\mathrm{i}t\tilde{\mathcal{Q}}(H)}E_{\mathcal{H}^{*}}=\mathop{\mbox{\rm s-lim}}_{\boldsymbol{\nu}\to\infty}\exp\left[-t\left(\boldsymbol{\nu}(\nabla^{*}\nabla-n)+\mathrm{i}H\right)\right],\quad t\geq 0,

where HH in the rhs is viewed as a multiplication operator MHM_{H} on Γ1,𝐌L2\Gamma_{1,{\bf M}}^{L^{2}}.

A similar statement and its proof are found in [14], but that proof was somewhat erroneous, and confusing in notations. Thus we will present another proof here. To complete the proof, we need some lemmas.

Lemma 4.3.

Let vv\in\mathcal{L} and vt:=et(A+iH)vv_{t}:=e^{-t\left(A+\mathrm{i}H\right)}v, t0t\geq 0. If vtdomAv_{t}\in\mathrm{dom}A and ddtvt|Avt0\frac{d}{dt}\left\langle v_{t}|Av_{t}\right\rangle\geq 0, then we have

AvtHvt,andvt|Avtα1H2vt2.\left\|Av_{t}\right\|\leq\left\|H\right\|\left\|v_{t}\right\|,\quad\text{and}\quad\left\langle v_{t}|Av_{t}\right\rangle\leq\alpha^{-1}\left\|H\right\|^{2}\left\|v_{t}\right\|^{2}.
Proof.

We see ddtvt|Avt=2(Avt|Avt+iHvt|Avt)\frac{d}{dt}\left\langle v_{t}|Av_{t}\right\rangle=-2\left(\left\langle Av_{t}|Av_{t}\right\rangle+\Re\left\langle\mathrm{i}Hv_{t}|Av_{t}\right\rangle\right) and |iHvt|Avt|HvtAvt\left|\left\langle\mathrm{i}Hv_{t}|Av_{t}\right\rangle\right|\leq\left\|H\right\|\left\|v_{t}\right\|\left\|Av_{t}\right\|. Hence ddtvt|Avt0\frac{d}{dt}\left\langle v_{t}|Av_{t}\right\rangle\geq 0 implies Avt2HvtAvt\left\|Av_{t}\right\|^{2}\leq\left\|H\right\|\left\|v_{t}\right\|\left\|Av_{t}\right\|, i.e. AvtHvt\left\|Av_{t}\right\|\leq\left\|H\right\|\left\|v_{t}\right\|. Since A2αAA^{2}\geq\alpha A, we have (αA)1/2vtHvt\|\left(\alpha A\right)^{1/2}v_{t}\|\leq\left\|H\right\|\left\|v_{t}\right\|. This is equivalent to vt|Avtα1H2vt2\left\langle v_{t}|Av_{t}\right\rangle\leq\alpha^{-1}\left\|H\right\|^{2}\left\|v_{t}\right\|^{2}.∎

Lemma 4.4.

Let v𝒦=kerAv\in\mathcal{K}=\ker A, and vt:=et(A+iH)vv_{t}:=e^{-t\left(A+\mathrm{i}H\right)}v, t0t\geq 0. Then for all t0t\geq 0, we have vtdomAv_{t}\in\mathrm{dom}A and

AvtHv,vt|AvtH2v2α,\left\|Av_{t}\right\|\leq\left\|H\right\|\left\|v\right\|,\quad\left\langle v_{t}|Av_{t}\right\rangle\leq\frac{\left\|H\right\|^{2}\left\|v\right\|^{2}}{\alpha}, (4.1)
E𝒦vtHvα.\left\|E_{\mathcal{K}}^{\perp}v_{t}\right\|\leq\frac{\left\|H\right\|\left\|v\right\|}{\alpha}. (4.2)
Proof.

The first and second inequalities are the direct consequences of Lemma 4.3. Recall α1AE𝒦\alpha^{-1}A\geq E_{\mathcal{K}}^{\perp}, then the third follows from

E𝒦vt2=vt|E𝒦vtα1vt|Avtα2H2v2.\left\|E_{\mathcal{K}}^{\perp}v_{t}\right\|^{2}=\left\langle v_{t}|E_{\mathcal{K}}^{\perp}v_{t}\right\rangle\leq\alpha^{-1}\left\langle v_{t}|Av_{t}\right\rangle\leq\alpha^{-2}\left\|H\right\|^{2}\left\|v\right\|^{2}.

Proof of Theorem 4.1: Substitute 𝝂A\boldsymbol{\nu}A for AA, and so 𝝂α\boldsymbol{\nu}\alpha for α\alpha in the above lemmas, and let v𝒦v\in\mathcal{K} and

vt:=vt(𝝂):=et(𝝂A+iH)vv_{t}:=v_{t}^{(\boldsymbol{\nu})}:=e^{-t\left(\boldsymbol{\nu}A+\mathrm{i}H\right)}v

We find

(ddt+iE𝒦HE𝒦)E𝒦vt(𝝂)=iE𝒦HE𝒦vt(𝝂),\left(\frac{d}{dt}+\mathrm{i}E_{\mathcal{K}}HE_{\mathcal{K}}\right)E_{\mathcal{K}}v_{t}^{(\boldsymbol{\nu})}=-\mathrm{i}E_{\mathcal{K}}HE_{\mathcal{K}}^{\perp}v_{t}^{(\boldsymbol{\nu})},

and hence

(ddt+iE𝒦HE𝒦)E𝒦vt(𝝂)=E𝒦HE𝒦vt(𝝂)HE𝒦vt(𝝂)\left\|\left(\frac{d}{dt}+\mathrm{i}E_{\mathcal{K}}HE_{\mathcal{K}}\right)E_{\mathcal{K}}v_{t}^{(\boldsymbol{\nu})}\right\|=\left\|E_{\mathcal{K}}HE_{\mathcal{K}}^{\perp}v_{t}^{(\boldsymbol{\nu})}\right\|\leq\left\|H\right\|\left\|E_{\mathcal{K}}^{\perp}v_{t}^{(\boldsymbol{\nu})}\right\|\ \
(4.2)HHv𝝂α=H2v𝝂α\leq_{\footnotesize(\ref{eq:Ev<Hv})}\left\|H\right\|\frac{\left\|H\right\|\left\|v\right\|}{\boldsymbol{\nu}\alpha}=\frac{\left\|H\right\|^{2}\left\|v\right\|}{\boldsymbol{\nu}\alpha}\ (4.3)

Thus we have

limn(ddt+iE𝒦HE𝒦)E𝒦vt(𝝂)=0.\lim_{n\to\infty}\left\|\left(\frac{d}{dt}+\mathrm{i}E_{\mathcal{K}}HE_{\mathcal{K}}\right)E_{\mathcal{K}}v_{t}^{(\boldsymbol{\nu})}\right\|=0.

This implies ut:=lim𝝂E𝒦vt(𝝂)u_{t}:=\lim_{\boldsymbol{\nu}\to\infty}E_{\mathcal{K}}v_{t}^{(\boldsymbol{\nu})} exists, and satisfies

(ddt+iE𝒦HE𝒦)ut=0,i.e.ut=eitE𝒦HE𝒦v.\left(\frac{d}{dt}+\mathrm{i}E_{\mathcal{K}}HE_{\mathcal{K}}\right)u_{t}=0,\quad\text{i.e.}\quad u_{t}=e^{-\mathrm{i}tE_{\mathcal{K}}HE_{\mathcal{K}}}v.

On the other hand, again by (4.2)

limnE𝒦vt(𝝂)lim𝝂Hv𝝂α=0.\lim_{n\to\infty}\left\|E_{\mathcal{K}}^{\perp}v_{t}^{(\boldsymbol{\nu})}\right\|\leq\lim_{\boldsymbol{\nu}\to\infty}\frac{\left\|H\right\|\left\|v\right\|}{\boldsymbol{\nu}\alpha}=0.

Hence we have

lim𝝂vt(𝝂)=lim𝝂E𝒦vt(𝝂)=ut.\lim_{\boldsymbol{\nu}\to\infty}v_{t}^{(\boldsymbol{\nu})}=\lim_{\boldsymbol{\nu}\to\infty}E_{\mathcal{K}}v_{t}^{(\boldsymbol{\nu})}=u_{t}.

5 Path integral representation

In this section, first we state a Feynman–Kac formula on a Riemannian manifold. Our main source is Güneysu [5].

Let (M,g)(M,g) be a complete Riemannian manifold. The Laplace-Beltrami operator on MM is denoted with

Δ=dd:C(M)C(M).\Delta=-\mathrm{d}^{*}\mathrm{d}:C^{\infty}(M)\rightarrow C^{\infty}(M).

The following fact is known (see e.g. [5], Theorem 2.24):

Theorem 5.1.

There exists a unique minimal positive fundamental solution

p:(0,)×M×M(0,),(t,x,y)pt(x,y)p:(0,\infty)\times M\times M\rightarrow(0,\infty),\quad(t,x,y)\mapsto p_{t}(x,y)

of the heat equation

th(t,x)=12Δxh(t,x).\frac{\partial}{\partial t}h(t,x)=\frac{1}{2}\Delta_{x}h(t,x).

The map pp is called the minimal heat kernel of M=(M,g)M=(M,g),

Definition 5.2.

Let XX be a continuous semi-martingale with values in MM in the time interval 𝐓=[t0,t1]{\bf T}=[t_{0},t_{1}] or [t0,)[t_{0},\infty). Fix a smooth principal bundle π:PM\pi:P\rightarrow M with structure group GG and the associated Lie algebra 𝔤\mathfrak{g}, and a connection 1-form α0𝐀1(P,𝔤)\alpha_{0}\in{\bf A}^{1}(P,\mathfrak{g}). A continuous semi-martingale UU on PP defined in the time interval 𝐓{\bf T} is called a horizontal lift of XX to PP (with respect to the connection α0\alpha_{0}), if π(U)=X\pi(U)=X and

α0(d¯U)=0,\int\alpha_{0}(\underline{\mathrm{d}}U)=0,

where d¯\underline{{\rm d}} denotes the Stratonovich integral.

Proposition 5.3.

([5], Theorem 2.15) There is a unique horizontal lift UU of XX to PP with U0=u0U_{0}=u_{0} \mathbb{P}-a.s.∎

Let EME\to M be a Hermitian vector bundle with a fixed Hermitian connection \nabla. Let π:P(E)M\pi:\mathrm{P}(E)\rightarrow M be the U(d)\mathrm{U}(d)-principal bundle corresponding to (E,(,)x)(E,(\bullet,\bullet)_{x}), that is,

P(E)=xM{u|u:dEx is an isometry}.\mathrm{P}(E)=\bigcup_{x\in M}\{u|u:\mathbb{C}^{d}\stackrel{{\scriptstyle\simeq}}{{\longrightarrow}}E_{x}\text{ is an isometry}\}.

The Hermitian connection \nabla on the vector bundle EME\to M is induced by a unique connection 1-form α0𝐀1(P(E),𝔤)\alpha_{0}\in{\bf A}^{1}(P(E),\mathfrak{g}) on PP, 𝔤:=𝔲(d)\mathfrak{g}:=\mathfrak{u}(d).

Let EEM×ME\boxtimes E^{*}\rightarrow M\times M be the external tensor product bundle corresponding to E,E, that is,

EE|(x,y)=ExEy=Hom(Ey,Ex).E\boxtimes E^{*}|_{(x,y)}=E_{x}\otimes E_{y}^{*}=\mathrm{Hom}(E_{y},E_{x}).
Proposition 5.4.

([5], Proposition and definition 2.17) Let XX be a continuous semi-martingale with values in MM in the time interval 𝐓=[t0,t1]{\bf T}=[t_{0},t_{1}] or [t0,)[t_{0},\infty). Let UU be a horizontal lift of XX to P(E)\mathrm{P}(E) w.r.t. the connection 1-form α0\alpha_{0}. Then the continuous adapted process given by

//X:=UU01:𝐓×ΩEE//^{X}:=UU_{0}^{-1}:{\bf T}\times\Omega\rightarrow E\boxtimes E^{*}

does not depend on the particular choice of the lift UU, and //X//^{X} is called the stochastic parallel transport in EE along X.X.

Let VLloc2(M,)V\in L_{{\rm loc}}^{2}(M,\mathbb{R}) and assume that VV is bounded from below, and is in the local Kato class. Define a self-adjoint operator

H(V):=12+VH(V):=\frac{1}{2}\nabla^{*}\nabla+V

on ΓL2(M,E)\Gamma_{L^{2}}(M,E), the Hilbert space of L2L^{2} sections of EE; Precisely, /2+V\nabla^{*}\nabla/2+V is shown to be essentially self-adjoint on the domain of smooth sections with compact support, and H(V)H(V) denotes the self-adjoint extension of it.

Proposition 5.5.

If MM is geodesically complete with Ricci curvature bounded from below and a positive injectivity radius, then there exists the Brownian bridge measures tx,y\mathbb{P}_{t}^{x,y} in a way that the expectation values 𝔼tx,y[]\mathbb{E}_{t}^{x,y}[\bullet] are a rigorous version of the conditional expectation values 𝔼[|Bt(x)=y]\mathbb{E}[\bullet|B_{t}(x)=y].∎

Theorem 5.6.

([5], Theorem 1.2) Let MM satisfy the conditions of Prop. 5.5, and BB a Brownian motion on 𝐌{\bf M} in the time interval [0,)[0,\infty). For any t>0t>0, define the section KtK_{t} of the bundle EEM×ME\boxtimes E^{*}\to M\times M, i.e., Kt(x,y)Hom(Ey,Ex)K_{t}(x,y)\in\mathrm{Hom}(E_{y},E_{x}) for all x,yMx,y\in M, by

Kt(x,y):=pt(x,y)𝔼tx,y[𝒱tB//tB,1],𝒱tB:=e0tV(Bs)ds,x,yM.K_{t}(x,y):=p_{t}(x,y)\mathbb{E}_{t}^{x,y}[\mathscr{V}_{t}^{B}//_{t}^{B,-1}],\quad\mathscr{V}_{t}^{B}:=e^{-\int_{0}^{t}V(B_{s})\mathrm{d}s},\quad x,y\in M. (5.1)

Then KtK_{t} defines an bounded integral kernel for the operator etH(V).e^{-tH(V)}. We write etH(V)(x,y):=Kt(x,y)e^{-tH(V)}(x,y):=K_{t}(x,y).∎

Here, VV is assumed to be real-valued. We want to extend this formula to the cases where VV is complex-valued. However, in that case H(V)H(V) is not self-adjoint, and the analysis appears to become far more difficult. The extension is easy only when VV is bounded, and its proof is similar to the case where VV is real-valued. We will consider this case in the following.

For each x,yMx,y\in M and 0t0<t10\leq t_{0}<t_{1}, let Cx,y([t0,t1],M)C_{x,y}([t_{0},t_{1}],M) denote the space of continuous functions 𝐱:[t0,t1]M{\bf x}:[t_{0},t_{1}]\to M such that 𝐱(t0)=x{\bf x}(t_{0})=x and 𝐱(t1)=y{\bf x}(t_{1})=y. Then we can choose Cx,y([0,t],M)C_{x,y}([0,t],M) (or more generally Cx,y([t0,t0+t],M)C_{x,y}([t_{0},t_{0}+t],M) as a natural sample space Ω\Omega of the probability measure tx,y\mathbb{P}_{t}^{x,y}. It follows from Theorem 5.6 that there exists a finite measure μ[0,t];x,y:=pt(x,y)tx,y\mu_{[0,t];x,y}:=p_{t}(x,y)\mathbb{P}_{t}^{x,y} on Cx,y([0,t],M)C_{x,y}([0,t],M) such that

etH(V)(x,y)=Cx,y([0,t],M)e0tV(𝐱t)ds//t𝐱,1dμ[0,t];x,y(𝐱).e^{-tH(V)}(x,y)=\int_{C_{x,y}([0,t],M)}e^{-\int_{0}^{t}V({\bf x}_{t})\mathrm{d}s}//_{t}^{{\bf x},-1}{\rm d}\mu_{[0,t];x,y}({\bf x}). (5.2)

In the following we use this form of path integral formula instead of the probabilistic form (5.1), mainly because the appearance of (5.2) is more similar to physicists’ naive path integral formulas than that of (5.1).

Recall the definitions and notations in previous sections, together with Assumption 3.1; Set M:=𝐌M:={\bf M}, and EE to be the Hermitian line bundle 𝒪1,𝐌\mathscr{O}_{1,{\bf M}} over 𝐌{\bf M}. Let \nabla be the Chern connection on 𝒪1,𝐌\mathscr{O}_{1,{\bf M}}.

For each 𝝂>0\boldsymbol{\nu}>0, consider the measure μ0,𝝂t;x,y\mu_{0,\boldsymbol{\nu}t;x,y} on Cx,y([0,𝝂t],𝐌)C_{x,y}([0,\boldsymbol{\nu}t],{\bf M}). Let 𝐱t(𝝂):=𝐱𝝂t{\bf x}_{t}^{(\boldsymbol{\nu})}:={\bf x}_{\boldsymbol{\nu}t}, and define the measure μ0,t;x,y(𝝂)\mu_{0,t;x,y}^{(\boldsymbol{\nu})} on Cx,y([0,t],𝐌)C_{x,y}([0,t],{\bf M}) by the time rescaling t𝝂tt\mapsto\boldsymbol{\nu}t:

dμ0,t;x,y(𝝂)(𝐱(𝝂)):=dμ0,𝝂t;x,y(𝐱).{\rm d}\mu_{0,t;x,y}^{(\boldsymbol{\nu})}({\bf x}^{(\boldsymbol{\nu})}):={\rm d}\mu_{0,\boldsymbol{\nu}t;x,y}({\bf x}).
Theorem 5.7.

Let 𝐌{\bf M} satisfy Assumption 3.1 and the conditions of Proposition 5.5 as a Riemannian manifold. Furthermore assume that the operator Δd′′\Delta_{{\rm d}^{\prime\prime}} on the line bundle 𝒪1,𝐌\mathscr{O}_{1,{\bf M}} has a spectral gap. Let HL(𝐌,)H\in L^{\infty}({\bf M},\mathbb{R}) (a “classical Hamiltonian”), and Ut:=eit𝒬(H)U_{t}:=e^{-\mathrm{i}t\mathcal{Q}(H)}, the corresponding “quantum time evolution”. (Note that Ut~=eit𝒬~(H)E=Eeit𝒬~(H)E.\widetilde{U_{t}}=e^{-\mathrm{i}t\tilde{\mathcal{Q}}(H)}E_{\mathcal{H}^{*}}=E_{\mathcal{H}^{*}}e^{-\mathrm{i}t\tilde{\mathcal{Q}}(H)}E_{\mathcal{H}^{*}}.) Then for any t>0t>0, the integral kernel of Ut~\widetilde{U_{t}} on 𝐌×𝐌{\bf M}\times{\bf M} is expressed as

Ut~(𝐩1,𝐩2)=lim𝝂e𝝂tnC𝐩1,𝐩2([0,t],𝐌)ei0tH(𝐱s)ds//t𝐱,1dμ0,t;𝐩1,𝐩2(𝝂)(𝐱),𝐩1,𝐩2𝐌,\widetilde{U_{t}}(\mathbf{p}_{1},\mathbf{p}_{2})=\lim_{\boldsymbol{\nu}\to\infty}e^{\boldsymbol{\nu}tn}\int_{C_{\mathbf{p}_{1},\mathbf{p}_{2}}([0,t],{\bf M})}e^{-\mathrm{i}\int_{0}^{t}H({\bf x}_{s}){\rm d}s}//_{t}^{{\bf x},-1}{\rm d}\mu_{0,t;\mathbf{p}_{1},\mathbf{p}_{2}}^{(\boldsymbol{\nu})}({\bf x}),\quad\mathbf{p}_{1},\mathbf{p}_{2}\in{\bf M},

where n=dim𝐌n=\dim_{\mathbb{R}}{\bf M}.

Proof.

Noticing Cor. 4.2 and

exp[t(𝝂(n)+iH)]=e𝝂tnexp[𝝂t(+i𝝂1H)],\exp\left[-t\left(\boldsymbol{\nu}(\nabla^{*}\nabla-n)+\mathrm{i}H\right)\right]=e^{\boldsymbol{\nu}tn}\exp\left[-\boldsymbol{\nu}t\left(\nabla^{*}\nabla+\mathrm{i}\boldsymbol{\nu}^{-1}H\right)\right],

let V:=i𝝂1HV:=\mathrm{i}\boldsymbol{\nu}^{-1}H. We see

//t𝐩(𝐱(𝝂))=//𝝂t𝐩(𝐱),0tH(𝐱s(𝝂))ds=0𝝂t𝝂1H(𝐱t)ds.//_{t}^{\mathbf{p}}({\bf x}^{(\boldsymbol{\nu})})=//_{\boldsymbol{\nu}t}^{\mathbf{p}}({\bf x}),\qquad\int_{0}^{t}H({\bf x}_{s}^{(\boldsymbol{\nu})}){\rm d}s=\int_{0}^{\boldsymbol{\nu}t}\boldsymbol{\nu}^{-1}H({\bf x}_{t})\mathrm{d}s.

Therefore, by (5.2),

exp[t(𝝂(n)+iH)](𝐩1,𝐩2)\displaystyle\exp\left[-t\left(\boldsymbol{\nu}(\nabla^{*}\nabla-n)+\mathrm{i}H\right)\right](\mathbf{p}_{1},\mathbf{p}_{2})
=e𝝂tnexp[𝝂t(+i𝝂1H)](𝐩1,𝐩2)\displaystyle\quad=e^{\boldsymbol{\nu}tn}\exp\left[-\boldsymbol{\nu}t\left(\nabla^{*}\nabla+\mathrm{i}\boldsymbol{\nu}^{-1}H\right)\right](\mathbf{p}_{1},\mathbf{p}_{2})
=e𝝂tnC𝐩1,𝐩2([0,𝝂t],M)e0𝝂ti𝝂1H(𝐱t)ds//𝝂t𝐩1,1(𝐱)dμ0,𝝂t;𝐩1,𝐩2(𝐱)\displaystyle\quad=e^{\boldsymbol{\nu}tn}\int_{C_{\mathbf{p}_{1},\mathbf{p}_{2}}([0,\boldsymbol{\nu}t],M)}e^{-\int_{0}^{\boldsymbol{\nu}t}\mathrm{i}\boldsymbol{\nu}^{-1}H({\bf x}_{t})\mathrm{d}s}//_{\boldsymbol{\nu}t}^{\mathbf{p}_{1},-1}({\bf x}){\rm d}\mu_{0,\boldsymbol{\nu}t;\mathbf{p}_{1},\mathbf{p}_{2}}({\bf x})
=e𝝂tnC𝐩1,𝐩2([0,t],M)ei0tH(𝐱s(𝝂))ds//t𝐩1,1(𝐱(𝝂))dμ0,t;𝐩1,𝐩2(𝝂)(𝐱(𝝂)),\displaystyle\quad=e^{\boldsymbol{\nu}tn}\int_{C_{\mathbf{p}_{1},\mathbf{p}_{2}}([0,t],M)}e^{-\mathrm{i}\int_{0}^{t}H({\bf x}_{s}^{(\boldsymbol{\nu})}){\rm d}s}//_{t}^{\mathbf{p}_{1},-1}({\bf x}^{(\boldsymbol{\nu})}){\rm d}\mu_{0,t;\mathbf{p}_{1},\mathbf{p}_{2}}^{(\boldsymbol{\nu})}({\bf x}^{(\boldsymbol{\nu})}),

and hence

eit𝒬~(H)(𝐩1,𝐩2)\displaystyle e^{-\mathrm{i}t\tilde{\mathcal{Q}}(H)}(\mathbf{p}_{1},\mathbf{p}_{2}) =lim𝝂exp[t(𝝂(n)+iH)](𝐩1,𝐩2)\displaystyle=\lim_{\boldsymbol{\nu}\to\infty}\exp\left[-t\left(\boldsymbol{\nu}(\nabla^{*}\nabla-n)+\mathrm{i}H\right)\right](\mathbf{p}_{1},\mathbf{p}_{2})
=lim𝝂e𝝂tnC𝐩1,𝐩2([0,t],M)ei0tH(𝐱s(𝝂))ds//t𝐩1,1(𝐱(𝝂))dμ0,t;𝐩1,𝐩2(𝝂)(𝐱(𝝂)).\displaystyle=\lim_{\boldsymbol{\nu}\to\infty}e^{\boldsymbol{\nu}tn}\int_{C_{\mathbf{p}_{1},\mathbf{p}_{2}}([0,t],M)}e^{-\mathrm{i}\int_{0}^{t}H({\bf x}_{s}^{(\boldsymbol{\nu})}){\rm d}s}//_{t}^{\mathbf{p}_{1},-1}({\bf x}^{(\boldsymbol{\nu})}){\rm d}\mu_{0,t;\mathbf{p}_{1},\mathbf{p}_{2}}^{(\boldsymbol{\nu})}({\bf x}^{(\boldsymbol{\nu})}).

Recall Lemma 2.2 and Eq.(3.1). Then we can define the integral kernel of Ut~\widetilde{U_{t}} on 𝕊(𝐌)×𝕊(𝐌)\mathbb{S}({\bf M})\times\mathbb{S}({\bf M}) (not on 𝐌×𝐌{\bf M}\times{\bf M}) by

Ut~(v1,v2):=v2|eit𝒬(H)v1=(Ut~(𝐩1,𝐩2)v2)(v1),vkran(𝐩k),k=1,2.\widetilde{U_{t}}(v_{1},v_{2}):=\bigl{\langle}v_{2}|e^{-\mathrm{i}t\mathcal{Q}(H)}v_{1}\bigr{\rangle}=\bigl{(}\widetilde{U_{t}}(\mathbf{p}_{1},\mathbf{p}_{2})v_{2}^{*}\bigr{)}(v_{1}),\quad v_{k}\in\mathrm{ran}(\mathbf{p}_{k}),\ k=1,2.

Similarly, the parallel transport //t𝐱,1Hom(𝒪1,𝐌,𝐩1,𝒪1,𝐌,𝐩2)=Hom(ran(𝐩1),ran(𝐩2))//_{t}^{{\bf x},-1}\in\mathrm{Hom}(\mathscr{O}_{1,{\bf M},\mathbf{p}_{1}},\mathscr{O}_{1,{\bf M},\mathbf{p}_{2}})=\mathrm{Hom}(\mathrm{ran}(\mathbf{p}_{1})^{*},\mathrm{ran}(\mathbf{p}_{2})^{*}) can also viewed as a function on ran(𝐩1)×ran(𝐩2)\mathrm{ran}(\mathbf{p}_{1})\times\mathrm{ran}(\mathbf{p}_{2}):

//t𝐱,1(v1,v2):=(//t𝐱,1v1)(v2),(v1,v2)ran(𝐩1)×ran(𝐩2).//_{t}^{{\bf x},-1}(v_{1},v_{2}):=\left(//_{t}^{{\bf x},-1}v_{1}^{*}\right)(v_{2}),\quad(v_{1},v_{2})\in\mathrm{ran}(\mathbf{p}_{1})\times\mathrm{ran}(\mathbf{p}_{2}).

If t1>t2t_{1}>t_{2}, let [t1,t2][t_{1},t_{2}] refer to the closed interval [t2,t1][t_{2},t_{1}], and let μt2,t1;𝐩2,𝐩1(𝝂)\mu_{t_{2},t_{1};\mathbf{p}_{2},\mathbf{p}_{1}}^{(\boldsymbol{\nu})} denote the same measure as μt1,t2;𝐩1,𝐩2(𝝂)\mu_{t_{1},t_{2};\mathbf{p}_{1},\mathbf{p}_{2}}^{(\boldsymbol{\nu})}.

Corollary 5.8.

For any t1,t2t_{1},t_{2}\in\mathbb{R}, v1,v2𝕊(𝐌)v_{1},v_{2}\in\mathbb{S}({\bf M}) with 𝐩k:=𝐩𝐫(vk)𝐌\mathbf{p}_{k}:={\bf pr}(v_{k})\in{\bf M} (k=1,2k=1,2),

v2|ei(t2t1)𝒬(H)v1=Ut2t1~(v1,v2)\displaystyle\bigl{\langle}v_{2}|e^{-\mathrm{i}(t_{2}-t_{1})\mathcal{Q}(H)}v_{1}\bigr{\rangle}=\widetilde{U_{t_{2}-t_{1}}}(v_{1},v_{2})
=lim𝝂en𝝂|t2t1|C𝐩1,𝐩2([t1,t2],𝐌)eit1t2H(𝐱s)ds//t2t1𝐱,1(v1,v2)dμt1,t2;𝐩1,𝐩2(𝝂)(𝐱).\displaystyle\quad=\lim_{\boldsymbol{\nu}\to\infty}e^{n\boldsymbol{\nu}|t_{2}-t_{1}|}\int_{C_{\mathbf{p}_{1},\mathbf{p}_{2}}([t_{1},t_{2}],{\bf M})}e^{-\mathrm{i}\int_{t_{1}}^{t_{2}}H({\bf x}_{s}){\rm d}s}//_{t_{2}-t_{1}}^{{\bf x},-1}(v_{1},v_{2}){\rm d}\mu_{t_{1},t_{2};\mathbf{p}_{1},\mathbf{p}_{2}}^{(\boldsymbol{\nu})}({\bf x}).

For k=1,,Nk=1,...,N, let t1<<tNt_{1}<\cdots<t_{N}, 𝐩k𝐌\mathbf{p}_{k}\in{\bf M}, vkran(𝐩k)v_{k}\in\mathrm{ran}(\mathbf{p}_{k}), vk=1\left\|v_{k}\right\|=1, and 𝐱k:[tk,tk+1]𝐌{\bf x}_{k}:[t_{k},t_{k+1}]\to{\bf M} (k=1,,N1k=1,...,N-1) be a path (of a Brownian motion) on 𝐌{\bf M} with 𝐱k(tk)=𝐩k{\bf x}_{k}(t_{k})=\mathbf{p}_{k}, 𝐱k(tk+1)=𝐩k+1{\bf x}_{k}(t_{k+1})=\mathbf{p}_{k+1}. Let 𝐱:[t1,tN]𝐌{\bf x}:[t_{1},t_{N}]\to{\bf M} be the concatenation of the paths 𝐱1,,𝐱N1{\bf x}_{1},...,{\bf x}_{N-1}. When N=3N=3, we have

//tN𝐱,1(v1,v3)=(//t3𝐱2,1//t2𝐱1,1)(v1,v3)=(//t3𝐱2,1//t2𝐱1,1v1)(v3)=v3|//t3𝐱2,1//t2𝐱1,1v1𝕊//_{t_{N}}^{{\bf x},-1}(v_{1},v_{3})=\left(//_{t_{3}}^{{\bf x}_{2},-1}//_{t_{2}}^{{\bf x}_{1},-1}\right)(v_{1},v_{3})=\left(//_{t_{3}}^{{\bf x}_{2},-1}//_{t_{2}}^{{\bf x}_{1},-1}v_{1}^{*}\right)(v_{3})=\bigl{\langle}v_{3}^{*}|//_{t_{3}}^{{\bf x}_{2},-1}//_{t_{2}}^{{\bf x}_{1},-1}v_{1}^{*}\bigr{\rangle}_{\mathbb{S}}
=v3|//t3𝐱2,1v2𝕊v2|//t2𝐱1,1v1𝕊=k=12//tk+1𝐱k,1(vk,vk+1),=\bigl{\langle}v_{3}^{*}|//_{t_{3}}^{{\bf x}_{2},-1}v_{2}^{*}\bigr{\rangle}_{\mathbb{S}}\bigl{\langle}v_{2}^{*}|//_{t_{2}}^{{\bf x}_{1},-1}v_{1}^{*}\bigr{\rangle}_{\mathbb{S}}=\prod_{k=1}^{2}//_{t_{k+1}}^{{\bf x}_{k},-1}(v_{k},v_{k+1}),

and in general

//tN𝐱,1(v1,vN)=(//tN𝐱N1,1//t2𝐱1,1)(v1,vN)=k=1N1//tk+1𝐱k,1(vk,vk+1)//_{t_{N}}^{{\bf x},-1}(v_{1},v_{N})=\left(//_{t_{N}}^{{\bf x}_{N-1},-1}\cdots//_{t_{2}}^{{\bf x}_{1},-1}\right)(v_{1},v_{N})=\prod_{k=1}^{N-1}//_{t_{k+1}}^{{\bf x}_{k},-1}(v_{k},v_{k+1})

Let 𝐩=(𝐩1,,𝐩N)𝐌N\vec{\mathbf{p}}=(\mathbf{p}_{1},...,\mathbf{p}_{N})\in{\bf M}^{N}, t=(t1,,tN)\vec{t}=(t_{1},...,t_{N}), 0t1<<tN0\leq t_{1}<\cdots<t_{N}.

Ct,𝐩:={𝐱C([t1,tN],𝐌)|𝐱(tj)=𝐩j,j=1,,N}C_{\vec{t},\vec{\mathbf{p}}}:=\left\{{\bf x}\in C([t_{1},t_{N}],{\bf M})|{\bf x}(t_{j})=\mathbf{p}_{j},\ j=1,...,N\right\}

Consider the measures μtj,tj+1;𝐩j,𝐩j+1(𝝂)\mu_{t_{j},t_{j+1};\mathbf{p}_{j},\mathbf{p}_{j+1}}^{(\boldsymbol{\nu})} on C𝐩j,𝐩j+1([tj,tj+1],𝐌)C_{\mathbf{p}_{j},\mathbf{p}_{j+1}}([t_{j},t_{j+1}],{\bf M}), and define the measure μt,𝐩(𝝂)\mu_{\vec{t},\vec{\mathbf{p}}}^{(\boldsymbol{\nu})} on Ct,𝐩C_{\vec{t},\vec{\mathbf{p}}} by

dμt,𝐩(𝝂)(𝐱):=j=1N1dμtj,tj+1;𝐩j,𝐩j+1(𝝂)(𝐱|[tj,tj+1]){\rm d}\mu_{\vec{t},\vec{\mathbf{p}}}^{(\boldsymbol{\nu})}({\bf x}):=\prod_{j=1}^{N-1}{\rm d}\mu_{t_{j},t_{j+1};\mathbf{p}_{j},\mathbf{p}_{j+1}}^{(\boldsymbol{\nu})}({\bf x}|_{[t_{j},t_{j+1}]})
Theorem 5.9.

For any 0t1<<tN0\leq t_{1}<\cdots<t_{N}, v1,,vN𝕊(𝐌)v_{1},...,v_{N}\in\mathbb{S}({\bf M}) with 𝐩k:=𝐩𝐫(vk)𝐌\mathbf{p}_{k}:={\bf pr}(v_{k})\in{\bf M} (k=1,,Nk=1,...,N), we have

k=1N1(U~tk+1tk(vk,vk+1))=lim𝝂en𝝂|tNt1|Ct,𝐩eit1tNH(𝐱s)ds//tN𝐱,1(v1,vN)dμt,𝐩(𝝂)(𝐱).\prod_{k=1}^{N-1}\left(\widetilde{U}_{t_{k+1}-t_{k}}(v_{k},v_{k+1})\right)=\lim_{\boldsymbol{\nu}\to\infty}e^{n\boldsymbol{\nu}|t_{N}-t_{1}|}\int_{C_{\vec{t},\vec{\mathbf{p}}}}e^{-\mathrm{i}\int_{t_{1}}^{t_{N}}H({\bf x}_{s}){\rm d}s}//_{t_{N}}^{{\bf x},-1}(v_{1},v_{N}){\rm d}\mu_{\vec{t},\vec{\mathbf{p}}}^{(\boldsymbol{\nu})}({\bf x}).
Remark 5.10.

We see

k=1N1(U~tk+1tk(vk,vk+1))\displaystyle\prod_{k=1}^{N-1}\left(\widetilde{U}_{t_{k+1}-t_{k}}(v_{k},v_{k+1})\right)
=vN|ei(tNtN1)𝒬(H)vN1v2|ei(t2t1)𝒬(H)v1\displaystyle=\bigl{\langle}v_{N}|e^{-\mathrm{i}(t_{N}-t_{N-1})\mathcal{Q}(H)}v_{N-1}\bigr{\rangle}\cdots\bigl{\langle}v_{2}|e^{-\mathrm{i}(t_{2}-t_{1})\mathcal{Q}(H)}v_{1}\bigr{\rangle}
=vN|ei(tNtN1)𝒬(H)𝐩N1ei(t3t2)𝒬(H)𝐩2ei(t2t1)𝒬(H)v1\displaystyle=\bigl{\langle}v_{N}|e^{-\mathrm{i}(t_{N}-t_{N-1})\mathcal{Q}(H)}\mathbf{p}_{N-1}\cdots e^{-\mathrm{i}(t_{3}-t_{2})\mathcal{Q}(H)}\mathbf{p}_{2}e^{-\mathrm{i}(t_{2}-t_{1})\mathcal{Q}(H)}v_{1}\bigr{\rangle}
=vN|eitN𝒬(H)𝐩N1,tN1𝐩2,t2eit1𝒬(H)v1\displaystyle=\bigl{\langle}v_{N}|e^{-\mathrm{i}t_{N}\mathcal{Q}(H)}\mathbf{p}_{N-1,t_{N-1}}\cdots\mathbf{p}_{2,t_{2}}e^{\mathrm{i}t_{1}\mathcal{Q}(H)}v_{1}\bigr{\rangle}

where 𝐩j,tj:=eitj𝒬~(H)𝐩jeitj𝒬~(H).\mathbf{p}_{j,t_{j}}:=e^{\mathrm{i}t_{j}\tilde{\mathcal{Q}}(H)}\mathbf{p}_{j}e^{-\mathrm{i}t_{j}\tilde{\mathcal{Q}}(H)}. Thus, this is invariant under the transformation vkeiθkvkv_{k}\mapsto e^{\mathrm{i}\theta_{k}}v_{k}, θk\theta_{k}\in\mathbb{R}, for k=2,,N1k=2,...,N-1. Furthermore, if v1=vNv_{1}=v_{N}, this equals

Tr(𝐩N1,tN1𝐩2,t2𝐩1,t1ei(tNt1)𝒬(H)){\rm Tr}\left(\mathbf{p}_{N-1,t_{N-1}}\cdots\mathbf{p}_{2,t_{2}}\mathbf{p}_{1,t_{1}}e^{-\mathrm{i}(t_{N}-t_{1})\mathcal{Q}(H)}\right)

and hence is invariant under the above transformation for all vkv_{k}, k=1,,Nk=1,...,N with θ1=θN\theta_{1}=\theta_{N}.

Proof.

By Corollary 5.8,

k=1N1U~tk+1tk(vk,vk+1)\displaystyle\prod_{k=1}^{N-1}\widetilde{U}_{t_{k+1}-t_{k}}(v_{k},v_{k+1})
=k=1N1lim𝝂e𝝂(tk+1tk)nC𝐩j,𝐩j+1([tk,tk+1],𝐌)eitktk+1H(𝐱s)ds\displaystyle\quad=\prod_{k=1}^{N-1}\lim_{\boldsymbol{\nu}\to\infty}e^{\boldsymbol{\nu}(t_{k+1}-t_{k})n}\int_{C_{\mathbf{p}_{j},\mathbf{p}_{j+1}}([t_{k},t_{k+1}],{\bf M})}e^{-\mathrm{i}\int_{t_{k}}^{t_{k+1}}H({\bf x}_{s}){\rm d}s}
×//tk+1𝐱k,1(vk,vk+1)dμtk,tk+1;𝐩k,𝐩k+1(𝝂)(𝐱k)\displaystyle\quad\quad\times//_{t_{k+1}}^{{\bf x}_{k},-1}(v_{k},v_{k+1}){\rm d}\mu_{t_{k},t_{k+1};\mathbf{p}_{k},\mathbf{p}_{k+1}}^{(\boldsymbol{\nu})}({\bf x}_{k})
=lim𝝂e𝝂(tNt1)n(k=1N1C𝐩j,𝐩j+1([tk,tk+1],𝐌)dμtk,tk+1;𝐩k,𝐩k+1(𝝂)(𝐱k))\displaystyle\quad=\lim_{\boldsymbol{\nu}\to\infty}e^{\boldsymbol{\nu}(t_{N}-t_{1})n}\left(\prod_{k=1}^{N-1}\int_{C_{\mathbf{p}_{j},\mathbf{p}_{j+1}}([t_{k},t_{k+1}],{\bf M})}{\rm d}\mu_{t_{k},t_{k+1};\mathbf{p}_{k},\mathbf{p}_{k+1}}^{(\boldsymbol{\nu})}({\bf x}_{k})\right)
×k=1N1eitktk+1H(𝐱s)ds//tk+1𝐱k,1(vk,vk+1)\displaystyle\quad\quad\times\prod_{k=1}^{N-1}e^{-\mathrm{i}\int_{t_{k}}^{t_{k+1}}H({\bf x}_{s}){\rm d}s}//_{t_{k+1}}^{{\bf x}_{k},-1}(v_{k},v_{k+1})
=lim𝝂e𝝂(tNt1)nCt,𝐩eit1tNH(𝐱s)ds//tN𝐱,1(v1,vN)dμt,𝐩(𝝂)(𝐱),\displaystyle\quad=\lim_{\boldsymbol{\nu}\to\infty}e^{\boldsymbol{\nu}(t_{N}-t_{1})n}\int_{C_{\vec{t},\vec{\mathbf{p}}}}e^{-\mathrm{i}\int_{t_{1}}^{t_{N}}H({\bf x}_{s}){\rm d}s}//_{t_{N}}^{{\bf x},-1}(v_{1},v_{N}){\rm d}\mu_{\vec{t},\vec{\mathbf{p}}}^{(\boldsymbol{\nu})}({\bf x}),

6 Representation of quantum probability

Recall (2.6), then we want a path integral formula for Tr(𝐩1,t1𝐩N,tN){\rm Tr}\left(\mathbf{p}_{1,t_{1}}\cdots\mathbf{p}_{N,t_{N}}\right) to represent the trace of product of some operators, which often has a physical meaning as a quantum probability or an expectation value of some physical quantity. Roughly speaking, if we set t1=tNt_{1}=t_{N} in Theorem 5.9, then we could get it. However, Theorem 5.9 is meaningful only when 0t1<<tN0\leq t_{1}<\cdots<t_{N}. If we want to extend Theorem 5.9 for any t1,,tNt_{1},...,t_{N}\in\mathbb{R}, we need to introduce a curve parameter 𝐭{\bf t} other than the time parameter tt.

Assume that H:𝐌H:{\bf M}\to\mathbb{R} is bounded and smooth. Consider the manifold 𝐌×{\bf M}\times\mathbb{R} with the natural Riemannian metric. Then 𝒪1,𝐌×\mathscr{O}_{1,{\bf M}}\times\mathbb{R} is a line bundle of 𝐌×{\bf M}\times\mathbb{R}. Define the connection (H)\nabla^{(H)} on 𝒪1,𝐌×\mathscr{O}_{1,{\bf M}}\times\mathbb{R} by

Xt(H)s:=Xs+iHts,X: a vector field on 𝐌,s: a section of 𝒪1,𝐌×,t:=t,\nabla_{X\oplus\partial_{t}}^{(H)}s:=\nabla_{X}s+\mathrm{i}H\partial_{t}s,\qquad X:\text{ a vector field on }{\bf M},\ s:\text{ a section of }\mathscr{O}_{1,{\bf M}}\times\mathbb{R},\ \partial_{t}:=\frac{\partial}{\partial t},

where \nabla is the Chern connection on 𝐌{\bf M}. For a path 𝐗{\bf X} on 𝐌×{\bf M}\times\mathbb{R}, let //(H)(𝐗)//^{(H)}({\bf X}) or //(H),𝐗//^{(H),{\bf X}} denote the parallel transport along 𝐗{\bf X} w.r.t. (H)\nabla^{(H)}.

Let t0t1t_{0}\leq t_{1}. For 𝐱C([t0,t1],𝐌){\bf x}\in C([t_{0},t_{1}],{\bf M}), define 𝐱~,𝐱~1C([0,1],𝐌×)\tilde{{\bf x}},\tilde{{\bf x}}^{-1}\in C([0,1],{\bf M}\times\mathbb{R}) by

𝐱~(𝐭):=(𝐱(t),t),t:=𝐭(t1t0)+t0,𝐭[0,1]\tilde{{\bf x}}({\bf t}):=\left({\bf x}\left(t\right),t\right),\quad t:={\bf t}\left(t_{1}-t_{0}\right)+t_{0},\ {\bf t}\in[0,1]
𝐱~1(𝐭):=𝐱~(1𝐭),𝐭[0,1].\tilde{{\bf x}}^{-1}({\bf t}):=\tilde{{\bf x}}(1-{\bf t}),\quad{\bf t}\in[0,1].

Notice that 𝐱~1=(𝐱~)1\tilde{{\bf x}}^{-1}=(\tilde{{\bf x}})^{-1} differs from (𝐱1)~\widetilde{({\bf x}^{-1})}. If 𝐱{\bf x} is piecewise smooth, we find

//t(H),𝐱~,1=ei0tH(𝐱s)ds//t𝐱,1.//_{t}^{(H),\tilde{{\bf x}},-1}=e^{-\mathrm{i}\int_{0}^{t}H({\bf x}_{s}){\rm d}s}//_{t}^{{\bf x},-1}. (6.1)

When 𝐱{\bf x} is a Brownian motion, this equation can be understood as a stochastic one.

For 𝐩0,𝐩1𝐌\mathbf{p}_{0},\mathbf{p}_{1}\in{\bf M} and t0,t1t_{0},t_{1}\in\mathbb{R}, let

C~t0,t1;𝐩0,𝐩1:={{𝐱~|𝐱C𝐩0,𝐩1([t0,t1],𝐌)}if t0t1{𝐱~1|𝐱C𝐩1,𝐩0([t1,t0],𝐌)}if t0>t1\tilde{C}_{t_{0},t_{1};\mathbf{p}_{0},\mathbf{p}_{1}}:=\begin{cases}\left\{\tilde{{\bf x}}|\ {\bf x}\in C_{\mathbf{p}_{0},\mathbf{p}_{1}}([t_{0},t_{1}],{\bf M})\right\}&\text{if }t_{0}\leq t_{1}\\ \left\{\tilde{{\bf x}}^{-1}|\ {\bf x}\in C_{\mathbf{p}_{1},\mathbf{p}_{0}}([t_{1},t_{0}],{\bf M})\right\}&\text{if }t_{0}>t_{1}\end{cases}

so that if 𝐗1C~t0,t1;𝐩0,𝐩1{\bf X}_{1}\in\tilde{C}_{t_{0},t_{1};\mathbf{p}_{0},\mathbf{p}_{1}} and 𝐗2C~t1,t2;𝐩1,𝐩2{\bf X}_{2}\in\tilde{C}_{t_{1},t_{2};\mathbf{p}_{1},\mathbf{p}_{2}}, then the concatenation 𝐗2𝐗1C([0,1],𝐌×){\bf X}_{2}\bullet{\bf X}_{1}\in C([0,1],{\bf M}\times\mathbb{R}) is defined:

(𝐗2𝐗1)(𝐭):={𝐗1(2𝐭)(0𝐭12)𝐗2(2𝐭1)(12<𝐭1)\left({\bf X}_{2}\bullet{\bf X}_{1}\right)({\bf t}):=\begin{cases}{\bf X}_{1}(2{\bf t})&\left(0\leq{\bf t}\leq\frac{1}{2}\right)\\ {\bf X}_{2}(2{\bf t}-1)&\left(\frac{1}{2}<{\bf t}\leq 1\right)\end{cases}

If t0t1t_{0}\leq t_{1} (resp. t0>t1t_{0}>t_{1}), the measure μt0,t1;𝐩1,𝐩2(𝝂)\mu_{t_{0},t_{1};\mathbf{p}_{1},\mathbf{p}_{2}}^{(\boldsymbol{\nu})} on C𝐩0,𝐩1([t0,t1],𝐌)C_{\mathbf{p}_{0},\mathbf{p}_{1}}([t_{0},t_{1}],{\bf M}) (resp. C𝐩1,𝐩0([t1,t0],𝐌)C_{\mathbf{p}_{1},\mathbf{p}_{0}}([t_{1},t_{0}],{\bf M})) induces the measure μ~t0,t1;𝐩0,𝐩1(𝝂)\tilde{\mu}_{t_{0},t_{1};\mathbf{p}_{0},\mathbf{p}_{1}}^{(\boldsymbol{\nu})} on C~t0,t1;𝐩0,𝐩1\tilde{C}_{t_{0},t_{1};\mathbf{p}_{0},\mathbf{p}_{1}}. Let 𝐩=(𝐩1,,𝐩N)𝐌N\vec{\mathbf{p}}=(\mathbf{p}_{1},...,\mathbf{p}_{N})\in{\bf M}^{N}, t=(t1,,tN)N\vec{t}=(t_{1},...,t_{N})\in\mathbb{R}^{N}. Define the space of loops Loopt,𝐩C([0,1],𝐌×){\rm Loop}_{\vec{t},\vec{\mathbf{p}}}\subset C([0,1],{\bf M}\times\mathbb{R}) by

Loopt,𝐩:={𝐗N𝐗1|𝐗kC~𝐩k,𝐩k+1;tk,tk+1,k=1,,N,𝐩N+1:=𝐩1,tN+1:=t1}{\rm Loop}_{\vec{t},\vec{\mathbf{p}}}:=\left\{{\bf X}_{N}\bullet\cdots\bullet{\bf X}_{1}|{\bf X}_{k}\in\tilde{C}_{\mathbf{p}_{k},\mathbf{p}_{k+1};t_{k},t_{k+1}},\ k=1,...,N,\ \mathbf{p}_{N+1}:=\mathbf{p}_{1},\ t_{N+1}:=t_{1}\right\}

where 𝐗N𝐗1:=𝐗N(𝐗N1(𝐗1)){\bf X}_{N}\bullet\cdots\bullet{\bf X}_{1}:={\bf X}_{N}\bullet\left({\bf X}_{N-1}\bullet\left(\cdots\bullet{\bf X}_{1}\right)\cdots\right), although the order of the concatenations is not important in the following; That is, we could alternatively define 𝐗N𝐗1{\bf X}_{N}\bullet\cdots\bullet{\bf X}_{1} by, say, ((𝐗N𝐗N1)𝐗1)\left(\cdots\left({\bf X}_{N}\bullet{\bf X}_{N-1}\right)\bullet\cdots\bullet{\bf X}_{1}\right).

Define the measure μt,𝐩(𝝂)\mu_{\vec{t},\vec{\mathbf{p}}}^{(\boldsymbol{\nu})} on Loop𝐩,t{\rm Loop}_{\vec{\mathbf{p}},\vec{t}} by

dμt,𝐩(𝝂)(𝐗):=k=1Ndμ~tk,tk+1;𝐩k,𝐩k+1(𝝂)(𝐗k),𝐗=𝐗N𝐗1Loop𝐩,t,{\rm d}\mu_{\vec{t},\vec{\mathbf{p}}}^{(\boldsymbol{\nu})}({\bf X}):=\prod_{k=1}^{N}{\rm d}\tilde{\mu}_{t_{k},t_{k+1};\mathbf{p}_{k},\mathbf{p}_{k+1}}^{(\boldsymbol{\nu})}({\bf X}_{k}),\quad{\bf X}={\bf X}_{N}\bullet\cdots\bullet{\bf X}_{1}\in{\rm Loop}_{\vec{\mathbf{p}},\vec{t}},

where 𝐩N+1:=𝐩1,tN+1:=t1\mathbf{p}_{N+1}:=\mathbf{p}_{1},\ t_{N+1}:=t_{1}. For 𝐗Loop𝐩,t{\bf X}\in{\rm Loop}_{\vec{\mathbf{p}},\vec{t}}, of course, //(H)(𝐗)U(1)//^{(H)}({\bf X})\in\mathrm{U}(1) refers to the (stochastic) holonomy along the loop 𝐗{\bf X}.

Lemma 6.1.

For any t1,t2t_{1},t_{2}\in\mathbb{R}, v1,v2𝕊(𝐌)v_{1},v_{2}\in\mathbb{S}({\bf M}) with 𝐩k:=𝐩𝐫(vk)𝐌\mathbf{p}_{k}:={\bf pr}(v_{k})\in{\bf M}, k=1,2k=1,2,

v2|ei(t2t1)𝒬(H)v1=lim𝝂en𝝂|t2t1|C~𝐩1,𝐩2([t1,t2],𝐌)//t2t1(H),𝐗,1(v1,v2)dμt1,t2;𝐩1,𝐩2(𝝂)(𝐗).\bigl{\langle}v_{2}|e^{-\mathrm{i}(t_{2}-t_{1})\mathcal{Q}(H)}v_{1}\bigr{\rangle}=\lim_{\boldsymbol{\nu}\to\infty}e^{n\boldsymbol{\nu}|t_{2}-t_{1}|}\int_{\tilde{C}_{\mathbf{p}_{1},\mathbf{p}_{2}}([t_{1},t_{2}],{\bf M})}//_{t_{2}-t_{1}}^{(H),{\bf X},-1}(v_{1},v_{2}){\rm d}\mu_{t_{1},t_{2};\mathbf{p}_{1},\mathbf{p}_{2}}^{(\boldsymbol{\nu})}({\bf X}).
Proof.

Directly follows from Corollary 5.8 and (6.1). ∎

Theorem 6.2.

Let 𝐩=(𝐩1,,𝐩N)𝐌N\vec{\mathbf{p}}=(\mathbf{p}_{1},...,\mathbf{p}_{N})\in{\bf M}^{N}, t=(t1,,tN)N\vec{t}=(t_{1},...,t_{N})\in\mathbb{R}^{N}. Then

Tr(𝐩1,t1𝐩N,tN)=lim𝝂enT𝝂Loop𝐩,t//(H)(𝐗)dμ𝐩,t(𝝂)(𝐗){\rm Tr}\left(\mathbf{p}_{1,t_{1}}\cdots\mathbf{p}_{N,t_{N}}\right)=\lim_{\boldsymbol{\nu}\to\infty}e^{nT\boldsymbol{\nu}}\int_{{\rm Loop}_{\vec{\mathbf{p}},\vec{t}}}//^{(H)}({\bf X}){\rm d}\mu_{\vec{\mathbf{p}},\vec{t}}^{(\boldsymbol{\nu})}({\bf X}) (6.2)

where T:=k=1N|tk+1tk|T:=\sum_{k=1}^{N}\left|t_{k+1}-t_{k}\right| (tN+1:=t1t_{N+1}:=t_{1}), 𝐩1,,𝐩N𝐌\mathbf{p}_{1},...,\mathbf{p}_{N}\in{\bf M} (k=1,,Nk=1,...,N).

Proof.

Let v1,,vN𝕊(𝐌)v_{1},...,v_{N}\in\mathbb{S}({\bf M}) with 𝐩k=𝐩𝐫(vk)𝐌\mathbf{p}_{k}={\bf pr}(v_{k})\in{\bf M}. Then we see

Tr(𝐩1,t1𝐩N,tN)¯=Tr(𝐩N,tN𝐩1,t1)=j=1Nvj+1|ei(tj+1tj)𝒬(H)vj\overline{{\rm Tr}\left(\mathbf{p}_{1,t_{1}}\cdots\mathbf{p}_{N,t_{N}}\right)}={\rm Tr}\left(\mathbf{p}_{N,t_{N}}\cdots\mathbf{p}_{1,t_{1}}\right)=\prod_{j=1}^{N}\bigl{\langle}v_{j+1}|e^{-\mathrm{i}(t_{j+1}-t_{j})\mathcal{Q}(H)}v_{j}\bigr{\rangle}

where vN+1:=v1v_{N+1}:=v_{1}, tN+1:=t1t_{N+1}:=t_{1}. Hence (6.2) follows from Lemma 6.1. ∎

Let (𝐌)\mathcal{B}({\bf M}) denote the family of Borel sets of 𝐌{\bf M}. Let

E0(S)\displaystyle E_{0}(S) :=𝒬(χS)=S𝐩dμ(𝐩),S(𝐌).\displaystyle:=\mathcal{Q}(\chi_{S})=\int_{S}\mathbf{p}\,{\rm d}\mu(\mathbf{p}),\qquad S\in\mathcal{B}({\bf M}).
Et(S)\displaystyle E_{t}(S) :=eit𝒬(H)E0(S)eit𝒬(H)=S𝐩tdμ(𝐩),t.\displaystyle:=e^{-\mathrm{i}t\mathcal{Q}(H)}E_{0}(S)e^{\mathrm{i}t\mathcal{Q}(H)}=\int_{S}\mathbf{p}_{t}\,{\rm d}\mu(\mathbf{p}),\quad t\in\mathbb{R}.

For each tt\in\mathbb{R}, Et()E_{t}(\bullet) is called a positive operator valued measure (POVM) on 𝐌{\bf M}. Let ρ\rho be a density operator on \mathcal{H}, i.e., ρ0\rho\geq 0, Trρ=1{\rm Tr}\rho=1, and let 0t1tN0\leq t_{1}\leq\cdots\leq t_{N}. Then the value

Pρ(t,S):=TrEtN(SN)Et1(S1)ρEt1(S1)EtN(SN),S:=(S1,,SN)P_{\rho}(\vec{t},\vec{S}):={\rm Tr}E_{t_{N}}(S_{N})\cdots E_{t_{1}}(S_{1})\rho E_{t_{1}}(S_{1})\cdots E_{t_{N}}(S_{N}),\quad\vec{S}:=(S_{1},...,S_{N})

is interpreted as the joint probability that under the condition that the state at time 0 is ρ\rho, the position in the phase space 𝐌{\bf M} is measured to be in S1S_{1} at time t1t_{1}, and then the position in the phase space 𝐌{\bf M} is measured to be in S2S_{2} at time t2t_{2}, etc. Of course, these measurements are somewhat “fuzzy” in that even if S,S(𝐌)S,S^{\prime}\in\mathcal{B}({\bf M}) satisfy SS=S\cap S^{\prime}=\emptyset, the probability Pρ((t,t),(S,S))P_{\rho}((t,t),(S,S^{\prime})) can be non-zero; Any error-free quantum measurement on 𝐌{\bf M} is impossible by the uncertainty principle.

Corollary 6.3.

Let ρ\rho be a density operator which have the representation ρ=𝒬(fρ)\rho=\mathcal{Q}(f_{\rho}) for some Borel function fρ:𝐌f_{\rho}:{\bf M}\to\mathbb{R}. Let

(F1,,F2N+1)\displaystyle(F_{1},...,F_{2N+1}) :=(χSN,,χS1,fρ,χS1,,χSN),\displaystyle:=(\chi_{S_{N}},...,\chi_{S_{1}},f_{\rho},\chi_{S_{1}},...,\chi_{S_{N}}),
(τ1,,τ2N+1)\displaystyle(\tau_{1},...,\tau_{2N+1}) :=(tN,,t1,t0,t1,,tN),t0:=0.\displaystyle:=(t_{N},...,t_{1},t_{0},t_{1},...,t_{N}),\qquad t_{0}:=0.

Then we have the path-integral representation of the quantum probability

Pρ(t,S)=lim𝝂e2ntN𝝂𝐌2N+1dμ2N+1(𝐩)Loop𝐩,τdμ𝐩,t(𝝂)(𝐗)//(H)(𝐗)j=12N+1Fj(𝐩j),P_{\rho}(\vec{t},\vec{S})=\lim_{\boldsymbol{\nu}\to\infty}e^{2nt_{N}\boldsymbol{\nu}}\int_{{\bf M}^{2N+1}}{\rm d}\mu^{2N+1}(\vec{\mathbf{p}})\int_{{\rm Loop}_{\vec{\mathbf{p}},\vec{\tau}}}{\rm d}\mu_{\vec{\mathbf{p}},\vec{t}}^{(\boldsymbol{\nu})}({\bf X})//^{(H)}({\bf X})\prod_{j=1}^{2N+1}F_{j}(\mathbf{p}_{j}),

where 𝐩=(𝐩1,,𝐩2N+1)\vec{\mathbf{p}}=(\mathbf{p}_{1},...,\mathbf{p}_{2N+1}), τ=(τ1,,τ2N+1)\vec{\tau}=(\tau_{1},...,\tau_{2N+1}), and

𝐌2N+1dμ2N+1(𝐩) denotes 𝐌dμ(𝐩1)𝐌dμ(𝐩2N+1).\int_{{\bf M}^{2N+1}}{\rm d}\mu^{2N+1}(\vec{\mathbf{p}})\ \text{ denotes }\int_{{\bf M}}{\rm d}\mu(\mathbf{p}_{1})\cdots\int_{{\bf M}}{\rm d}\mu(\mathbf{p}_{2N+1}).
Proof.

By Theorem 6.2 and Eq.(2.6), we find

Pρ(t,S)\displaystyle P_{\rho}(\vec{t},\vec{S}) =Tr𝒬tN(χSN)𝒬t1(χS1)𝒬(fρ)𝒬t1(χS1)𝒬tN(χSN)\displaystyle={\rm Tr}\mathcal{Q}_{t_{N}}(\chi_{S_{N}})\cdots\mathcal{Q}_{t_{1}}(\chi_{S_{1}})\mathcal{Q}(f_{\rho})\mathcal{Q}_{t_{1}}(\chi_{S_{1}})\cdots\mathcal{Q}_{t_{N}}(\chi_{S_{N}})
=𝐌dμ(𝐩1)𝐌dμ(𝐩2N+1)Tr(𝐩1,t1𝐩N,tN)j=1NFj(𝐩j)\displaystyle=\int_{{\bf M}}{\rm d}\mu(\mathbf{p}_{1})\cdots\int_{{\bf M}}{\rm d}\mu(\mathbf{p}_{2N+1}){\rm Tr}\left(\mathbf{p}_{1,t_{1}}\cdots\mathbf{p}_{N,t_{N}}\right)\prod_{j=1}^{N}F_{j}(\mathbf{p}_{j})
=𝐌dμ(𝐩1)𝐌dμ(𝐩2N+1)[j=12N+1Fj(𝐩j)]lim𝝂enT𝝂Loop𝐩,t//(H)(𝐗)dμ𝐩,t(𝝂)(𝐗),\displaystyle=\int_{{\bf M}}{\rm d}\mu(\mathbf{p}_{1})\cdots\int_{{\bf M}}{\rm d}\mu(\mathbf{p}_{2N+1})\left[\prod_{j=1}^{2N+1}F_{j}(\mathbf{p}_{j})\right]\lim_{\boldsymbol{\nu}\to\infty}e^{nT\boldsymbol{\nu}}\int_{{\rm Loop}_{\vec{\mathbf{p}},\vec{t}}}//^{(H)}({\bf X}){\rm d}\mu_{\vec{\mathbf{p}},\vec{t}}^{(\boldsymbol{\nu})}({\bf X}),

with

T:=k=12N|τk+1τk|=2j=0N1(tk+1tk)=2tN.T:=\sum_{k=1}^{2N}|\tau_{k+1}-\tau_{k}|=2\sum_{j=0}^{N-1}(t_{k+1}-t_{k})=2t_{N}.

Remark 6.4.

Note that there exist many density operators ρ\rho which do not have the representation ρ=𝒬(fρ)\rho=\mathcal{Q}(f_{\rho}). However, the set of density operators of the form 𝒬(f)\mathcal{Q}(f) is dense in the space of density operators, which has a metric induced by the trace norm, and hence any density operator can be approximated by the operator of the form 𝒬(f)\mathcal{Q}(f).

Remark 6.5.

We also emphasize that the formulas for quantum joint probabilities as Corollary 6.3 will not be formulated in the imaginary-time path integral; There will be no direct relation between the quantum joint probability (with real-time evolution) and the imaginary-time path integral.

7 Example: Glauber coherent states

Consider the case where 𝐌{\bf M}\subset\mathbb{P}\mathcal{H} is homeomorphic to n2n\mathbb{C}^{n}\cong\mathbb{R}^{2n}. Let ψ:n×\psi:\mathbb{C}^{n}\to\mathcal{H}^{\times} be a holomorphism such that 𝐩𝐫ψ(n)=𝐌{\bf pr}\circ\psi(\mathbb{C}^{n})={\bf M}, so that 𝐩𝐫ψ:n𝐌{\bf pr}\circ\psi:\mathbb{C}^{n}\to{\bf M} is a single global coordinate chart of 𝐌{\bf M}. Without loss of generality, assume ψ(0)|ψ(z)=1\left\langle\psi(0)|\psi(z)\right\rangle=1 for all znz\in\mathbb{C}^{n}.

The line bundle 𝒪1,𝐌\mathscr{O}_{1,{\bf M}} over 𝐌{\bf M}, which is topologically trivial, i.e. 𝐌×\cong{\bf M}\times\mathbb{C}, has the Hermitian metric |pt\left\langle\cdot|\cdot\right\rangle_{{\rm pt}} given by f1|f2pt(𝐩𝐫(v)):=v2f1(v)¯f2(v)\left\langle f_{1}|f_{2}\right\rangle_{{\rm pt}}({\bf pr}(v)):=\left\|v\right\|^{-2}\overline{f_{1}\left(v\right)}f_{2}\left(v\right), where v𝐩𝐫1(𝐌),f1,f2Γ1,𝐌.v\in{\bf pr}^{-1}({\bf M}),\ f_{1},f_{2}\in\Gamma_{1,{\bf M}}. A global Kähler potential 𝐡:𝐌{\bf h}:{\bf M}\to\mathbb{R} is given by

𝐡(z)=1h(z)=ψ(z)2.{\bf h}(z)=\frac{1}{h(z)}=\left\|\psi(z)\right\|^{2}.

The corresponding global symplectic form is ω:=id′′dlog𝐡\omega:=-\mathrm{i}\,{\rm d}^{\prime\prime}{\rm d}^{\prime}\log{\bf h}, and the Riemannian metric is g(X,Y):=ω(X,JY)g(X,Y):=\omega(X,JY). Let 𝐞:=ψ(0){\bf e}:=\psi(0) then 𝐞{\bf e}^{*} is a global holomorphic section of 𝒪1,𝐌\mathscr{O}_{1,{\bf M}}; Precisely, if we identify the fiber of 𝒪1,𝐌\mathscr{O}_{1,{\bf M}} over 𝐩𝐌\mathbf{p}\in{\bf M} with ran(𝐩)\mathrm{ran}(\mathbf{p})^{*}, the value 𝐞(𝐩){\bf e}^{*}(\mathbf{p}) of the section 𝐞Γ1,𝐌{\bf e}^{*}\in\Gamma_{1,{\bf M}} at 𝐩𝐌\mathbf{p}\in{\bf M} is the function ran(𝐩)\mathrm{ran}(\mathbf{p})\to\mathbb{C}, v𝐞|vv\mapsto\left\langle{\bf e}|v\right\rangle. The normalization of 𝐞{\bf e}^{*} is 𝐞¯:=h1/2𝐞\underline{{\bf e}}^{*}:=h^{-1/2}{\bf e}^{*}; Precisely, the value 𝐞¯(𝐩)\underline{{\bf e}}^{*}(\mathbf{p}) of the section 𝐞¯Γ1,𝐌\underline{{\bf e}}^{*}\in\Gamma_{1,{\bf M}} at 𝐩𝐌\mathbf{p}\in{\bf M} is the function

ran(𝐩),ζψ(z)¯ζ,zn,ψ(z)ran(𝐩),ζ.\mathrm{ran}(\mathbf{p})\to\mathbb{C},\qquad\zeta\underline{\psi(z)}\mapsto\zeta,\qquad z\in\mathbb{C}^{n},\ \psi(z)\in\mathrm{ran}(\mathbf{p}),\ \zeta\in\mathbb{C}.

Here the Chern connection form w.r.t. the holomorphic frame 𝐞{\bf e}^{*} is a \mathbb{C}-valued 1-form on n\mathbb{C}^{n} defined by θ:=dlogh=dlog𝐡ι\theta:={\rm d}^{\prime}\log h=-{\rm d}^{\prime}\log{\bf h}_{\iota}. The Chern connection is globally defined by

(f𝐞):=(df)𝐞+fθ𝐞,fΓ0,𝐌.\nabla\left(f{\bf e}^{*}\right):=\left({\rm d}f\right)\otimes{\bf e}^{*}+f\theta\otimes{\bf e}^{*},\qquad f\in\Gamma_{0,{\bf M}}^{\infty}.

The Chern connection form θnor\theta_{{\rm nor}} on n\mathbb{C}^{n} w.r.t. the normalized frame 𝐞¯\underline{{\bf e}}^{*} is defined by (3.3).

Let C:[0,1]nC:[0,1]\to\mathbb{C}^{n} be a piecewise smooth curve, so that 𝐱:=𝐩𝐫ψC{\bf x}:={\bf pr}\circ\psi\circ C is a piecewise smooth curve on 𝐌{\bf M}. If CC is a loop, i.e. C(0)=C(1)C(0)=C(1), the parallel transport (holonomy) //1𝐱//_{1}^{{\bf x}} becomes a scalar:

//𝐱//1𝐱=eiCθnorU(1).//^{{\bf x}}\equiv//_{1}^{{\bf x}}=e^{\mathrm{i}\int_{C}\theta_{{\rm nor}}}\in\mathrm{U}(1).

Assume the conditions of Corollary 6.3. Let 𝐗Loopt,𝐩{\bf X}\in{\rm Loop}_{\vec{t},\vec{\mathbf{p}}}, with 𝐗(𝐭)=(𝐱(𝐭),γ(𝐭)){\bf X}({\bf t})=({\bf x}({\bf t}),\gamma({\bf t})) (𝐭[0,1]{\bf t}\in[0,1]) where 𝐱:[0,1]𝐌{\bf x}:[0,1]\to{\bf M}, γ:[0,1]\gamma:[0,1]\to\mathbb{R}. Consider the manifold 𝐌×{\bf M}\times\mathbb{R} as in Sec. 6, and let H:𝐌H:{\bf M}\to\mathbb{R} be a bounded smooth Hamiltonian. The connection on the line bundle 𝒪1,𝐌×\mathscr{O}_{1,{\bf M}}\times\mathbb{R} over 𝐌×{\bf M}\times\mathbb{R} defined in Sec. 6 is explicitly given by the global \mathbb{R}-valued 1-form

ΘH:=θnor+Hdt,\Theta_{H}:=\theta_{{\rm nor}}+H{\rm d}t,

on n×\mathbb{C}^{n}\times\mathbb{R}, w.r.t. the normalized frame of 𝒪1,𝐌×\mathscr{O}_{1,{\bf M}}\times\mathbb{R} , and hence the holonomy along 𝐗{\bf X} becomes

//(H)(𝐗)=exp(i𝐗^ΘH),𝐗^(𝐭):=(ψ1(𝐱(𝐭)),γ(𝐭))n×.//^{(H)}({\bf X})=\exp\left(\mathrm{i}\int_{\hat{{\bf X}}}\Theta_{H}\right),\qquad\hat{{\bf X}}({\bf t}):=(\psi^{-1}({\bf x}({\bf t})),\gamma({\bf t}))\in\mathbb{C}^{n}\times\mathbb{R}.

Let Loop^t,𝐩:={𝐗^|𝐗Loopt,𝐩}\widehat{{\rm Loop}}_{\vec{t},\vec{\mathbf{p}}}:=\left\{\hat{{\bf X}}|{\bf X}\in{\rm Loop}_{\vec{t},\vec{\mathbf{p}}}\right\}. The measure μ𝐩,t(𝝂)\mu_{\vec{\mathbf{p}},\vec{t}}^{(\boldsymbol{\nu})} on Loopt,𝐩{\rm Loop}_{\vec{t},\vec{\mathbf{p}}} induces the measure μ^𝐩,t(𝝂)\hat{\mu}_{\vec{\mathbf{p}},\vec{t}}^{(\boldsymbol{\nu})} on Loop^t,𝐩\widehat{{\rm Loop}}_{\vec{t},\vec{\mathbf{p}}} by the bijection 𝐗𝐗^{\bf X}\mapsto\hat{{\bf X}}. Now the quantum joint probability formula in Corollary 6.3 is rewritten in a somewhat more familiar and intuitive form:

Theorem 7.1.

Under the conditions of Corollary 6.3,

Pρ(t,S)=lim𝝂e2ntN𝝂𝐌2N+1dμ2N+1(𝐩)Loop^𝐩,τdμ^𝐩,t(𝝂)(𝐗^)ei𝐗^ΘHj=12N+1Fj(𝐩j).P_{\rho}(\vec{t},\vec{S})=\lim_{\boldsymbol{\nu}\to\infty}e^{2nt_{N}\boldsymbol{\nu}}\int_{{\bf M}^{2N+1}}{\rm d}\mu^{2N+1}(\vec{\mathbf{p}})\int_{\widehat{{\rm Loop}}_{\vec{\mathbf{p}},\vec{\tau}}}{\rm d}\hat{\mu}_{\vec{\mathbf{p}},\vec{t}}^{(\boldsymbol{\nu})}(\hat{{\bf X}})e^{\mathrm{i}\int_{\hat{{\bf X}}}\Theta_{H}}\prod_{j=1}^{2N+1}F_{j}(\mathbf{p}_{j}).
Example 7.2.

Next let us consider the most basic but important case. Let n=1n=1. Let a,aa^{*},a be usual creation/annihilation operators on a Hilbert space \mathcal{H}, which satisfy [a,a]=1[a,a^{*}]=1, and assume that {a,a}\{a,a^{*}\} is irreducible on \mathcal{H}. Let v0𝕊()v_{0}\in\mathbb{S}(\mathcal{H}) be a “vacuum state, ” i.e. av0=0av_{0}=0. Set

ψ(z)=ezav0,z\psi(z)=e^{za^{*}}v_{0},\qquad z\in\mathbb{C}

then the corresponding Kähler potential, the symplectic form, and the Riemannian metric are calculated as

𝐡(z)=ψ(z)2=e|z|2,{\bf h}(z)=\left\|\psi(z)\right\|^{2}=e^{|z|^{2}},
ω=id′′dlog𝐡=idzdz¯=2dxdy,g=2dxdy.\omega=-\mathrm{i}\,{\rm d}^{\prime\prime}{\rm d}^{\prime}\log{\bf h}=\mathrm{i}\,{\rm d}z\wedge{\rm d}\overline{z}=2{\rm d}x\wedge{\rm d}y,\qquad g=2{\rm d}x\otimes{\rm d}y.

Thus 𝐌{\bf M}\cong\mathbb{C} as Kähler manifolds; The volume measure on 𝐌{\bf M} is the usual Lebesgue measure dxdy{\rm d}x{\rm d}y on 2\mathbb{C}\cong\mathbb{R}^{2} times 2. Note that the normalized vector |z:=ψ(z)¯:=e|z|2/2ψ(z)|z\rangle:=\underline{\psi(z)}:=e^{-|z|^{2}/2}\psi(z) is nothing other than the coherent state in the usual sense (i.e. the Glauber coherent state). We can also check ψ(0)|ψ(z)=1\left\langle\psi(0)|\psi(z)\right\rangle=1. The overcompleteness relation

1π|zz|dxdy=I,z=x+iy\frac{1}{\pi}\int_{\mathbb{C}}|z\rangle\langle z|{\rm d}x{\rm d}y=I,\qquad z=x+\mathrm{i}y

implies that we should take the measure μ\mu on 𝐌{\bf M} as dμ(z):=1πdxdy{\rm d}\mu(z):=\frac{1}{\pi}{\rm d}x{\rm d}y. Now the BS quantization of f:𝐌f:{\bf M}\to\mathbb{R} is given by

𝒬(f)=f^(z)|zz|dμ(z),where f^(z):=f(𝐩𝐫(ψ(z)))=f(|zz|).\mathcal{Q}(f)=\int_{\mathbb{C}}\hat{f}(z)|z\rangle\langle z|{\rm d}\mu(z),\quad\text{where }\hat{f}(z):=f({\bf pr}(\psi(z)))=f(|z\rangle\langle z|).

The r.h.s. is often called the Glauber–Sudarshan representation of the operator of l.h.s.

𝐞=ψ(0)=v0{\bf e}^{*}=\psi(0)^{*}=v_{0}^{*} is a holomorphic section of 𝒪1,𝐌\mathscr{O}_{1,{\bf M}}. By (3.3) we find

θnor=ydxxdy,z=x+iy.\theta_{{\rm nor}}=y{\rm d}x-x{\rm d}y,\qquad z=x+\mathrm{i}y.

Let C:[0,1]C:[0,1]\to\mathbb{C} be a piecewise smooth curve, with 𝐱:=𝐩𝐫ψC{\bf x}:={\bf pr}\circ\psi\circ C. Then we find that the operation of the parallel transport //t𝐱//_{t}^{{\bf x}} on 𝒪1,𝐌\mathscr{O}_{1,{\bf M}} is explicitly written as

//t𝐱C(0)|=eiC[0,t]θnorC(t)|,//_{t}^{{\bf x}}\langle C(0)|=e^{\mathrm{i}\int_{C\upharpoonright[0,t]}\theta_{{\rm nor}}}\langle C(t)|,

where we used the bra notation z|:=ψ(z)¯\langle z|:=\underline{\psi(z)^{*}}, zz\in\mathbb{C}. Let ΘH:=θnor+Hdt\Theta_{H}:=\theta_{{\rm nor}}+H{\rm d}t, then the quantum joint probability w.r.t. the time evolution generated by the quantum Hamiltonian 𝒬(H)\mathcal{Q}(H) is given by Theorem 7.1.

Note that the paths occurring in the (stochastic) line integral in this formula are closed, and hence even if we substitute an arbitrary θnor:=θnor+α\theta_{{\rm nor}}^{\prime}:=\theta_{{\rm nor}}+\alpha with dα=0{\rm d}\alpha=0 for θnor\theta_{{\rm nor}}, we get the same probability. To fix the 1-form θnor\theta_{{\rm nor}} is analogous to a gauge fixing in physics; This suggests that Theorem 7.1 realizes a somewhat “gauge-invariant” formulation of quantization. However, of course, different gauge fixings of a classical system can lead to different quantizations, where the difference is experimentally observable, in general. Thus we can only say that Theorem 7.1 may reduce the “gauge-dependence” of the notion of quantizations.

8 Toward the geometric path integral quantizations

The results in the previous sections are given in the situation where the BS quantization 𝒬(f)\mathcal{Q}(f) is considered only when the function ff is bounded, and so 𝒬(f)\mathcal{Q}(f) becomes a bounded operator. The various difficulties in dealing with unbounded operators put obstacles in the generalizations of those results, and so it may be very difficult to formulate a general theory. (One can see such difficulty also from the case of deformation quantization; The theory of deformation quantization was presented by [2] in 1978. The theory achieved a great success in the algebraic level, when Kontsevich [8] proved its generality, that there exists a deformation quantization for every Poisson manifold. However, the theories to realize the deformation quantizations in terms of the operators in the Hilbert spaces (e.g. the strict deformation quantization of Rieffel [10, 11]) seem to remain incomplete even on the problems concerning only bounded operators.)

Nevertheless, we conjecture that the results of the previous sections can be generalized for most situations which are “physically relevant.” To clarify this conjecture, we consider the notion of “geometric path integral quantization” in this section.

Let 𝐌{\bf M} be a complete Kähler manifold, and assume the condition of Proposition 5.5 as a Riemannian manifold. (Forget the assumption in Sec. 2 that 𝐌{\bf M} is a submanifold of some projective space \mathbb{P}\mathcal{H}.) Physically, we interpret 𝐌{\bf M} as a classical-mechanical phase space, whose symplectic form is the Kähler form ω\omega. Recall the notion of prequantization bundle in geometric quantization [13].

Definition 8.1.

A symplectic manifold (𝐌,ω)({\bf M},\omega) is prequantizable when there exists a Hermitian line bundle, called a prequantization bundle, π:L𝐌\pi:L\rightarrow{\bf M} with connection \nabla, whose curvature form Θ\Theta is proportional to the symplectic 2-form, Θ=iω/\Theta=-\mathrm{i}\omega/\hbar. (Cf. Eq.(3.5))

Note: In this paper, we set =1\hbar=1, and recall that Θ\Theta is defined to be a 𝔲(1)=i\mathfrak{u}(1)=\mathrm{i}\mathbb{R}-valued 2-form here.

For quantization, we must assume that 𝐌{\bf M} is prequantizable. Moreover we assume that the prequantization bundle L𝐌L\to{\bf M} is holomorphic.

Conjecture 8.2.

Assume that H:𝐌H:{\bf M}\to\mathbb{R} satisfy some adequate conditions as a classical-mechanical Hamiltonian. Then Cor. 6.3 can be generalized for such HH.

Remark 8.3.

In the typical cases, H(x)H(x) is smooth, bounded from below, and increases in a polynomial order as x\left\|x\right\|\to\infty. Hence we could take the above “adequate conditions” as such conditions. However also note the important cases such as the Hamiltonian of a hydrogen atom, which is not bounded from below.

Note that since a Brownian motion on 𝐌{\bf M} is well-defined, our path integral is also well-defined for each fixed 𝝂\boldsymbol{\nu}. The above assertion says that the probability Pρ(t,S)P_{\rho}(\vec{t},\vec{S}) calculated via our path integral (with the limit 𝝂\boldsymbol{\nu}\to\infty) coincides with the value calculated via BS quantization. However, since 𝐌{\bf M} is not assumed to be a subset of \mathbb{P}\mathcal{H} here, BS quantization is not defined yet, and so the above assertion is still too vague. We will explain further this point in the following.

Consider the Hilbert space 𝒦:=ΓholL2ΓL2\mathcal{K}:=\Gamma_{{\rm hol}}^{L^{2}}\subset\Gamma^{L^{2}}, where ΓL2\Gamma^{L^{2}} is the space of L2L^{2} sections of BB, and ΓholL2\Gamma_{{\rm hol}}^{L^{2}} is the closed subspace consisting of holomorphic sections. Here we consider 𝒦\mathcal{K} as the quantum state space, following the method of holomorphic quantization.

Let KAK_{A} be the integral kernel of an operator AA on ΓL2\Gamma^{L^{2}}, i.e.,

(As)(x1)=𝐌KA(x1,x2)s(x2)dx2,sΓL2,x1𝐌.\left(As\right)(x_{1})=\int_{{\bf M}}K_{A}(x_{1},x_{2})s(x_{2})\,{\rm d}x_{2},\quad s\in\Gamma^{L^{2}},\ x_{1}\in{\bf M}.

where dx2{\rm d}x_{2} denotes the integral w.r.t. the volume form vol{\rm vol} on 𝐌{\bf M}. KAK_{A} is a map such that KA(x1,x2)Hom(Lx2,Lx1)K_{A}(x_{1},x_{2})\in\mathrm{Hom}(L_{x_{2}},L_{x_{1}}) for all x1,x2𝐌x_{1},x_{2}\in{\bf M}, where LxL_{x} is the fiber of the line bundle LL at x𝐌x\in{\bf M}; Equivalently, KAK_{A} is a section of the external tensor product bundle LL𝐌×𝐌L\boxtimes L^{*}\to{\bf M}\times{\bf M}.

Let E𝒦E_{\mathcal{K}} denote the orthogonal projection from ΓL2\Gamma^{L^{2}} onto 𝒦\mathcal{K}. For each x𝐌x\in{\bf M}, define vx𝕊(𝒦)ΓL2v_{x}\in\mathbb{S}(\mathcal{K})\subset\Gamma^{L^{2}} by

vx(x):=CxKE𝒦(x,x),x,x𝐌,Cx>0.v_{x}(x^{\prime}):=C_{x}K_{E_{\mathcal{K}}}(x^{\prime},x),\quad x,x^{\prime}\in{\bf M},\ C_{x}>0.

Define P:𝐌𝒦P:{\bf M}\to\mathbb{P}\mathcal{K} by

P(x):=|vxvx|,x𝐌.P(x):=|v_{x}\rangle\langle v_{x}|,\qquad x\in{\bf M}.

Assume that this map PP is an embedding of the Kähler manifold 𝐌{\bf M} into 𝒦\mathbb{P}\mathcal{K}, viewed as a possibly infinite-dimensional Kähler manifold. This assumption says that 𝐌{\bf M} may be identified with its range P(𝐌)𝒦P({\bf M})\subset\mathbb{P}\mathcal{K}, in other words, that 𝐌{\bf M} can be seen as a submanifold of the projective space 𝒦\mathbb{P}\mathcal{K}. In this situation the BS quantization can be defined for 𝐌{\bf M}, and so the meaning of Conjecture 8.2 becomes clearer.

Acknowledgement.

I am grateful to Koichi Arashi for several useful remarks on a preliminary version of the manuscript.

References

  • [1] Werner Ballmann. Lectures on Kähler Manifolds. ESI Lectures in Mathematics and Physics. European Mathematical Society, 2006. http://people.mpim-bonn.mpg.de/hwbllmnn/archiv/kaehler0609.pdf.
  • [2] F. Bayen, M. Flato, C. Fronsdal, A. Lichnerowicz, and D. Sternheimer. Deformation theory and quantization I. Ann. Phys. (NY), 111:61–110, 1978.
  • [3] I. Daubechies and J. R. Klauder. Quantum-mechanical path integrals with Wiener measure for all polynomial Hamiltonians II. J. Math. Phys., 26(9):2239–2256, 1985.
  • [4] R. P. Feynman and A. Hibbs. Quantum Mechanics and Path Integrals. McGraw-Hill, New York, 1965.
  • [5] Batu Güneysu. On the Feynman-Kac formula for Schrödinger semigroups on vector bundles. PhD thesis, Rheinischen Friedrich-Wilhelms-Universität Bonn, 2010. https://hdl.handle.net/20.500.11811/4970.
  • [6] A. A. Kirillov. Lectures on the Orbit Method. A.M.S., Providence, 2004.
  • [7] Shoshichi Kobayashi. Differential geometry of complex vector bundles. Iwanami Shoten, and Princeton University Press, 1987. This is 2013 re-typesetting of the book 1987: http://mathsoc.jp/publication/PublMSJ/PDF/Vol15.pdf.
  • [8] M. Kontsevich. Deformation quantization of Poisson manifolds. Lett. Math. Phys., 66(3):157–216, 2003. q-alg/9709040.
  • [9] M. Reed and B. Simon. Fourier Analysis, Self-adjointness. Methods of Modern Mathematical Physics II. Academic Press, San Diego, 1975.
  • [10] M. A. Rieffel. Deformation quantization of heisenberg manifolds. Comm. Math. Phys., 122:531–562, 1989.
  • [11] Marc A. Rieffel. Deformation guantization for Actions of RdR^{d}. American Mathematical Society, Providence, 1993.
  • [12] B. Simon. The classical limit of quantum partition functions. Commun. Math. Phys., 71:247–276, 1980.
  • [13] N. M. J. Woodhouse. Geometric Quantization. Clarendon Press, Oxford, 2nd edition edition, 1992.
  • [14] H. Yamashita. Phase-space path integral and Brownian motion. J. Math. Phys., 52:022101, 2011.
  • [15] H. Yamashita. Glauber-Sudarshan-type quantizations and their path integral representations for compact Lie groups. arXiv:1811.08844, 2018.