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

Spectral action and heat kernel trace for Ricci flat manifolds from stochastic flow over second quantized L2L^{2}-differential forms.

Sita Gakkhar111Sitanshu Gakkhar, email. The first author would like to thank Debashish Goswami for helpful discussions on qsde’s of Evans-Hudson type, and also Branimir Ćaćić for discussions on Dirac operators, Matilde Marcolli,
Abstract

A quantum stochastic differential equation (qsde) on Fock space over L2L^{2} differential 11-forms is given from the small “time” flow of which the trace of the connection Laplacian heat kernel for the spinor endomorphism bundle can be computed over any compact Ricci-flat Riemannian manifold. The existence of the stochastic flow is established by adapting the construction from [14]. When the manifold supports a parallel spinor – Ricci-flatness is a required integrability condition for parallel spinors, the trace of Dirac Laplacian heat kernel of the spinor bundle can be recovered. For 44-manifolds, this corresponds to the spectral action, and realizes Einstein-Hilbert action as a stochastic flow.

1 Introduction

This article attempts to formalize the idea that Einstein-Hilbert action for a Riemannian manifold can be viewed as arising from random fluctuations acting on the spinor bundle. The model of random fluctuations is provided by a stochastic flow generated by the Dirac Laplacian. The connection between Dirac Laplacian, noncommutative geometry and gravity is well established (see [15] and references therein for an account), the new contribution here is the probabilistic perspective.

This realization of spectral action as a stochastic flow is suggested by explicit computations of spectral action which yield Brownian bridge integrals along with the observation that boson Fock space can be viewed as a Wiener space, with a preliminary exploration of the perspective put forward in [9] where covariant quantum diffusions on almost-commutative spectral triples are considered. This article treats canonical spectral triples, computing explicitly the structure matrix for the diffusion generated by the endomorphism connection Laplacian acting on the algebra of functions over any compact Riemannian manifold. In absence of a natural action with respect to which the generator is covariant, the standard constructions using Picard iterates (see [14], and also [1]) care adapted to show the existence of the solution. From this flow, when there exists a parallel spinor, spectral action can be evaluated.

Some remarks on notation. By Riemannian (M,g)(M,g), we mean a Riemannian manifold MM with metric gg. The connection on the tangent bundle of MM, TMTM, is always the Levi-Civita connection unless specified otherwise. When clear from context, the same symbol is used of the connection \nabla on a Hermitian or Riemannian bundle EE and the dual connection on dual bundle EE^{*}. After fixing a local orthonormal frame about any pMp\in M, (Xi)idimM(X_{i})_{i\in\dim M}, Xi\nabla_{X_{i}} will be used interchangeably with i\nabla_{i}. For local coordinates (xi)(x_{i}) about any pMp\in M, i\partial_{i} will denote the coordinate vector fields xi\frac{\partial}{\partial x_{i}}. [n][n] is the set {i,in}\{i\in{\mathds{N}},i\leq n\} where {\mathds{N}}, with convention that 00\not\in{\mathds{N}}. [n:m][n:m] denotes the set {n,n+1m}\{n,n+1\dots m\}. The linear span we mean finite linear span denoted by LinSpan(V):={i[k]αiai:αi𝕂,aiV}\operatorname{{LinSpan}}({\textsf{V}}):=\{\sum_{i\in[k]}\alpha_{i}a_{i}:\alpha_{i}\in\mathbb{K},a_{i}\in{\textsf{V}}\} where 𝕂=,\mathbb{K}=\mathds{R},\mathbb{C} is clear from context; if 𝕂=\mathbb{K}=\mathds{R} then the subscript is dropped. Throughout Γ(H)\operatorname{{\Gamma}}(H) denotes the symmetric (boson) Fock space over the any space HH, while E(H)\textsc{E}(H) denotes the exponential vectors given by E(v)=n=0(n!)1/2vn\textsc{E}(v)=\oplus_{n=0}^{\infty}(n!)^{-1/2}v^{\otimes n} for vHv\in H. From section 3 onwards, we make use of Einstein notation. All calculation prior to section 3 are using at the center of normal coordinates where the metric is identity, for clarity, the explicit summations are used.

1.1 Organization and overview

In the remainder of this section, we introduce the background on stochastic flows and spectral action, and delve into motivating ideas. In section 2, the structure matrix for connection Laplacian is computed following the standard prescription. The noise space turns out to be L2L^{2} differential forms and the flow lives on the Fock space over the differential forms. The section concludes by writing the structure matrix and the qsde for the stochastic flow in the coordinate free quantum stochastic calculus notation; the relevant material from quantum stochastic differential equations an appendix reviewing quantum stochastic differential equations is included at the end (appendix B). Necessary bounds for controlling growth of Laplacian powers are established in section 3. The existence of flows for the derived qsde is established in section 4 by providing estimates that can be plugged into the standard theory.

1.2 Spectral action and stochastic flows

The canonical spectral triple for Riemannian (M,g)(M,g) which carries a spin structure is the data (C(M),L2(S),DM)(C^{\infty}(M),L^{2}(S),D_{M}), L2(S)L^{2}(S) being the Hilbert space of square integrable sections of a spinor bundle SMS\to M, and DMD_{M} the Dirac operator associated to the lift of Levi-Civita connection to the spinor bundle[16, pg 67]. We will take SS to be any spinor bundle associated to TMTM and DMD_{M} to be the Atiyah-Singer operator \not{D} (see [11, ex II.5.9]), then we have that 2=+κ/4\not{D}^{2}=\mathop{}\!\mathbin{\bigtriangleup}+\kappa/4 where \mathop{}\!\mathbin{\bigtriangleup} is the connection Laplacian for the connection on SS and κ\kappa is the scalar curvature[11, Thm II.8.8].

The bosonic spectral action is the linear funtional SbM:=Trf(DM/Λ)S_{b}^{M}:=\operatorname{{Tr}}f(D_{M}/\Lambda) for a choice of test function ff which we take as ex2e^{-x^{2}} and Λ\Lambda a cutoff parameter\cites[§ 5.1]ncg_standard_model[§ 7.1]suijlekom_ncg, so that the parameter tt of the Dirac heat semigroup etDM2e^{-tD_{M}^{2}} will be taken to satisfy t=Λ2t=\Lambda^{-2}. From the asymptotic expansion limΛSbM\lim_{\Lambda\to\infty}S_{b}^{M} for a Riemannian spin 44-manifold MM, the Einstein-Hilbert action, SEHS_{EH}, can be recovered\cites[§ 5.3]ncg_standard_model[§ 8.3]suijlekom_ncg.

Since DM,DM2D_{M},D_{M}^{2} are self-adjoint on L2(S)L^{2}(S), suppose (ϕi,j)(\phi_{i,j}) is an basis of orthonormal eigensections with eigenvalues λi,j2\lambda_{i,j}^{2} for DM2D^{2}_{M} where j[ni]j\in[n_{i}] runs over the multiplicity nin_{i} of eigenvalue λi,j2:=λi2\lambda^{2}_{i,j}:=\lambda^{2}_{i} Define the state Φz=1/N(z)i:λizj[ni]ϕij\Phi^{z}=1/N(z)\sum_{i:\lambda_{i}\leq z}\sum_{j\in[n_{i}]}\phi_{i}^{j} with normalization N(z):=i:λizniN(z):=\sqrt{\sum_{i:\lambda_{i}\leq z}n_{i}} then

limzΦz,etDM2ΦzN(z)\displaystyle\lim_{z\to\infty}{\langle\Phi^{z},e^{-tD_{M}^{2}}\Phi^{z}\rangle}N(z) =i:λizjniϕi,j,etDM2ϕi,j=limzΦz,etDM2Φz=TretDM2\displaystyle=\sum_{i:\lambda_{i}\leq z}\sum_{j\in n_{i}}{\langle\phi_{i,j},e^{-tD_{M}^{2}}\phi_{i,j}\rangle}=\lim_{z\to\infty}{\langle\Phi^{z},e^{-tD_{M}^{2}}\Phi^{z}\rangle}=\operatorname{{Tr}}e^{-tD_{M}^{2}}

Therefore, the bosonic spectral action can be approximated by expectation of small time expectation of et(DM)2e^{-t(D_{M})^{2}} in state Φz\Phi^{z} for zz large:

SEHlimt0TretDM2=limt0limzΦz,etDM2ΦzN(z)S_{EH}\approx\lim_{\sqrt{t}\to 0}\operatorname{{Tr}}e^{-tD_{M}^{2}}=\lim_{\sqrt{t}\to 0}\lim_{z\to\infty}{\langle\Phi^{z},e^{-tD_{M}^{2}}\Phi^{z}\rangle}N(z)

This motivates the interest in evaluating Φz,etDM2Φz{\langle\Phi^{z},e^{-tD_{M}^{2}}\Phi^{z}\rangle}. The approach we take is that of a quantum stochastic dilation associated to the heat semigroup, etDM2e^{-tD^{2}_{M}}, which is a quantum dynamical semigroup by [9].

The issue is complicated by the fact that for the existence of Evans-Husdon flow, the generator must annihilate identity, that is, the semigroup must be conservative. The Dirac laplacian acting on endomorphisms by composition does not satisfy this. We instead have to work with the endomorphism connection and the associated endomorphism connection laplacian and endomorphism Dirac laplacian and then extract the spinor bundle Dirac laplacian from it (see [9] more detailed discussion on this). Because of Ricci flatness, 2=\not{D}^{2}=\mathop{}\!\mathbin{\bigtriangleup}, the semigroup of interest will be etEnde^{-t\mathop{}\!\mathbin{\bigtriangleup}^{\operatorname{{End}}}} for the endomorphism connection laplacian End\mathop{}\!\mathbin{\bigtriangleup}^{\operatorname{{End}}}.

Remark 1.1.

If the spinor bundle could be replaced by the Clifford bundle, then the associated Dirac laplacian does generate a flow of Evans-Hudson type and the endomorphism trick is not needed.

A quantum stochastic dilation of the Evans-Hudson type222see appendix B for a brief review of quantum stochastic processes and integrals, and [14, 13] for detailed standard theory for the quantum dynamical semigroup ete^{-t{\mathcal{L}}} on the C*-algebra 𝒜:=𝒞(M)End(L2(S)){\mathcal{A}}:={\mathcal{C}}(M)\subset\operatorname{{End}}(L^{2}(S)) is a family of *-homomorphisms jt:𝒜𝒜′′(Γ(+,k0))j_{t}:{\mathcal{A}}\to{\mathcal{A}}^{\prime\prime}\otimes{\mathcal{B}}(\operatorname{{\Gamma}}(\mathds{R}_{+},k_{0})) where k0k_{0} is a Hilbert space, called the noise (or multiplicity) Hilbert space that is constructed from the generator -{\mathcal{L}}, such that the following holds

  • There exists an ultra-weak dense *-subalgebra 𝒜0𝒜{\mathcal{A}}_{0}\subset{\mathcal{A}} such that the map-valued process Jt:𝒜E(L2(+,k0))𝒜′′Γ(L2(+,k0))J_{t}:{\mathcal{A}}\otimes\textsc{E}(L^{2}(\mathds{R}_{+},k_{0}))\to{\mathcal{A}}^{\prime\prime}\otimes\operatorname{{\Gamma}}(L^{2}(\mathds{R}_{+},k_{0})) with Jt(xE(f))u:=jt(x)(uE(f))J_{t}(x\otimes\textsc{E}(f))u:=j_{t}(x)(u\textsc{E}(f)) for x𝒜,uH:=L2(S)x\in{\mathcal{A}},u\in{\textsf{H}}:=L^{2}(S), fL2(+,k0)f\in L^{2}(\mathds{R}_{+},k_{0}), satisfying the qsde:

    dJt=Jt(aδ(dt)+aδ(dt)+Λσ(dt)+I(dt)),J0=1\displaystyle dJ_{t}=J_{t}\circ(a_{\delta}(dt)+a_{\delta}^{\dagger}(dt)+\Lambda_{\sigma}(dt)+I_{\mathcal{L}}(dt)),J_{0}={\textbf{1}} (1)

    on 𝒜0E(L2(+,k0)){\mathcal{A}}_{0}\otimes\textsc{E}(L^{2}(\mathds{R}_{+},k_{0})) where δ:𝒜0𝒜k0,σ:𝒜0𝒜(k0)\delta:{\mathcal{A}}_{0}\to{\mathcal{A}}\otimes k_{0},\sigma:{\mathcal{A}}_{0}\to{\mathcal{A}}\otimes{\mathcal{B}}(k_{0}) are linear maps, called the structure maps for the qsde, derived from the generator {\mathcal{L}} for etDM2e^{-tD_{M}^{2}}, 𝒜0Dom(){\mathcal{A}}_{0}\subset\operatorname{{Dom}}({\mathcal{L}}), and aδ,aδ,Λσ,Ia_{\delta},a^{\dagger}_{\delta},\Lambda_{\sigma},I_{\mathcal{L}} the fundamental processes with respect to which stochastic integral is defined.

  • For all u,vH,x𝒜u,v\in{\textsf{H}},x\in{\mathcal{A}},

    vE(0),jt(x)E(0)=v,etu\displaystyle{\langle v\textsc{E}(0),j_{t}(x)\textsc{E}(0)\rangle}={\langle v,e^{-t{\mathcal{L}}}u\rangle} (2)

Therefore, Φz,etΦz{\langle\Phi^{z},e^{-t{\mathcal{L}}}\Phi_{z}\rangle} is realized as operator algebraic expectation of jtj_{t} with respect vacuum state |ΦzE(0)L2(S)Γ(L2(+,k0)){|\Phi^{z}\textsc{E}(0)\rangle}\in L^{2}(S)\otimes\operatorname{{\Gamma}}(L^{2}(\mathds{R}_{+},k_{0})), where jtj_{t} is obtained from the flow JtJ_{t} for the Evans-Hudson qsde (equation 1). Schemes for solving for Evans-Hudson qsde, for example, using Picard iterates, provide a way to algorithmically construct the flow.

1.3 Dirac Laplacian, endomorphism connection and parallel spinors

To start we note the following sign conventions of the Laplacians. Primarily the signs are fixed so the Laplace-Beltrami operator has non-negative spectrum, and signs on all other Laplacians cascade from there. On Riemannian (M,g)(M,g), MM compact, without boundary, Trg\operatorname{{Tr}}_{g} denotes the trace of a covariant tensor taken after identifying with a contravariant tensor via the metric gg, Trg(h):=gijhij\operatorname{{Tr}}_{g}(h):=g^{ij}h_{ij}. Note that trace on contravariant tensor, e.g. vector fields, is simply the sum. For XΓ(TM)X\in\Gamma(TM), div(X)=Tr(X)\text{div}(X)=\operatorname{{Tr}}(\nabla X). The Laplace-Beltrami operator is taken as the operator with non-negative spectrum, that is, div()=Tr(,)-\text{div}(\nabla)=-\operatorname{{Tr}}(\nabla_{\cdot,\cdot}), where ,\nabla_{\cdot,\cdot} is the second invariant derivative V,W2:=VWVW:Γ(E)Γ(E)\nabla^{2}_{V,W}:=\nabla_{V}\nabla_{W}-\nabla_{\nabla_{V}W}:\Gamma(E)\to\Gamma(E). The connection Laplacian is \nabla^{*}\nabla where \nabla^{*} is adjoint of the connection :Γ(E)Γ(E)TM\nabla:\Gamma(E)\to\Gamma(E)\otimes T^{*}M. Equivalently, =Tr(,)\nabla^{*}\nabla=-\operatorname{{Tr}}(\nabla_{\cdot,\cdot}). Further, =gijij\mathop{}\!\mathbin{\bigtriangleup}=-g^{ij}\nabla_{i}\nabla_{j}.

Let \nabla be any connection on the vector bundle EME\to M, (M,g)(M,g) a compact Riemannian manifold. The connection Laplacian at pMp\in M, =\mathop{}\!\mathbin{\bigtriangleup}=\nabla^{*}\nabla in local coordinates (ei)(e_{i}) is given by =(iiiiei\mathop{}\!\mathbin{\bigtriangleup}=-(\sum_{i}\nabla_{i}\nabla_{i}-\nabla_{\nabla_{i}e_{i}}). To evaluate ϕ\mathop{}\!\mathbin{\bigtriangleup}\phi at any pMp\in M and ϕΓ(E)\phi\in\Gamma(E), we will use Riemann normal coordinates centered at pp so iej\nabla_{i}e_{j} vanish, yielding ϕ(p)=iiiϕ(p)\mathop{}\!\mathbin{\bigtriangleup}\phi(p)=-\sum_{i}\nabla_{i}\nabla_{i}\phi(p).

The endomorphism connection End\nabla^{\operatorname{{End}}} on the bundle End(E)=EE\operatorname{{End}}(E)=E\otimes E^{*} associated to a connection \nabla on the Hermitian vector bundle EE over the Riemannian manifold MM is such that for XTM,XEnd=X1+1~XX\in TM,\nabla_{X}^{\operatorname{{End}}}=\nabla_{X}\otimes 1+1\otimes\widetilde{\nabla}_{X}, where ~\widetilde{\nabla} is the dual connection on EE^{*}. The endomorphism Laplacian is defined as usual: at pMp\in M in Riemann normal coordinates centered at pp (denoting ~~,~\widetilde{\nabla}^{*}\widetilde{\nabla},\widetilde{\nabla} by ,\mathop{}\!\mathbin{\bigtriangleup},\nabla again),

End=iiEndiEnd=(1+2iii+1)\mathop{}\!\mathbin{\bigtriangleup}^{\operatorname{{End}}}=-\sum_{i}\nabla_{i}^{\operatorname{{End}}}\nabla_{i}^{\operatorname{{End}}}=-\left(\mathop{}\!\mathbin{\bigtriangleup}\otimes 1+2\sum_{i}\nabla_{i}\otimes\nabla_{i}+1\otimes\mathop{}\!\mathbin{\bigtriangleup}\right)

Note that as EEE\otimes E^{*} is balanced over 𝒞(M){\mathcal{C}}(M), the action of 𝒞(M){\mathcal{C}}(M) on End(E)\operatorname{{End}}(E) can be written as f1End=i(fhi)hif\cdot 1_{\operatorname{{End}}}=\sum_{i}(f\cdot h_{i})\otimes h_{i}^{*}; this convention is used for all computation with Laplacian expressed in this tensor form. It’s also very useful to note that in any local coordinates , End\nabla^{\operatorname{{End}}} acts by commutator: if over chart UU, the connection has potential AA, =d+A\nabla=d+A, then for a local orthonormal frame (μi)(\mu_{i}) and dual frame (μj)(\mu^{j}), Endijσjiμiμj=ij(dσji)μiμj+jk[σAAσ]jkμkμj\nabla^{\operatorname{{End}}}\sum_{ij}\sigma^{i}_{j}\mu_{i}\otimes\mu^{j}=\sum_{ij}(d\sigma_{j}^{i})\mu_{i}\otimes\mu^{j}+\sum_{jk}[\sigma A-A\sigma]_{jk}\mu_{k}\otimes\mu^{j}. In particular, since 1End{\textbf{1}}_{\operatorname{{End}}} is given by the identity matrix locally, it follows (see [9]) that End(1End)=0\nabla^{\operatorname{{End}}}({\textbf{1}}_{\operatorname{{End}}})=0. This implies that again in normal Riemann coordinates centered at pp yields that for any f𝒞(M)f\in{\mathcal{C}}^{\infty}(M),

End(f1End)(p)\displaystyle\mathop{}\!\mathbin{\bigtriangleup}^{\operatorname{{End}}}(f{\textbf{1}}_{\operatorname{{End}}})(p) =iiEndiEnd(f1End)(p)=iiEnd(if1End)=iiif1End(p)\displaystyle=-\sum_{i}\nabla^{\operatorname{{End}}}_{i}\nabla^{\operatorname{{End}}}_{i}(f\cdot{\textbf{1}}_{\operatorname{{End}}})(p)=\sum_{i}\nabla_{i}^{\operatorname{{End}}}\left(\partial_{i}f{\textbf{1}}_{\operatorname{{End}}}\right)=-\sum_{i}\partial_{i}\partial_{i}f\cdot{\textbf{1}}_{\operatorname{{End}}}(p) (3)
Proposition 1.2.

For f𝒞(M)f\in{\mathcal{C}}^{\infty}(M), End(E)(f1End)=M(f)1End\mathop{}\!\mathbin{\bigtriangleup}^{\operatorname{{End}}(E)}(f\cdot{\textbf{1}}_{\operatorname{{End}}})=\mathop{}\!\mathbin{\bigtriangleup}^{M}(f)\cdot{\textbf{1}}_{\operatorname{{End}}}.

Proof.

Let Γijk\Gamma^{k}_{ij} be the Christoffel symbols for Levi-Civita connection, then in local coordinates about xMx\in M, M(f)=ijgij(x)(ijkΓijkk)f\mathop{}\!\mathbin{\bigtriangleup}^{M}(f)=-\sum_{ij}g^{ij}(x)(\partial_{i}\partial_{j}-\sum_{k}\Gamma^{k}_{ij}\partial_{k})f (see, for instance, [2, pg 66]) and for the endomorphism Laplacian,

End(f1)\displaystyle\mathop{}\!\mathbin{\bigtriangleup}^{\operatorname{{End}}}(f\cdot 1) =ijgij(x)(iEndjEndkΓijkkEnd)(f1End)\displaystyle=-\sum_{ij}g^{ij}(x)(\nabla^{\operatorname{{End}}}_{i}\nabla^{\operatorname{{End}}}_{j}-\sum_{k}\Gamma^{k}_{ij}\nabla^{\operatorname{{End}}}_{k})(f\cdot{\textbf{1}}_{\operatorname{{End}}})
=(ijgij(x)(ijkΓijkk)f)1End=M(f)1\displaystyle=-\left(\sum_{ij}g^{ij}(x)(\partial_{i}\partial_{j}-\sum_{k}\Gamma^{k}_{ij}\partial_{k})f\right)\cdot{\textbf{1}}_{\operatorname{{End}}}=\mathop{}\!\mathbin{\bigtriangleup}^{M}(f){\textbf{1}}

where we used End(1End)=0,XEnd(f)1End=X(f)1End\nabla^{\operatorname{{End}}}({\textbf{1}}_{\operatorname{{End}}})=0,\nabla_{X}^{\operatorname{{End}}}(f)\cdot{\textbf{1}}_{\operatorname{{End}}}=X(f)\cdot{\textbf{1}}_{\operatorname{{End}}}. ∎

Now for any constant cc, Tr(et(+c))\operatorname{{Tr}}(e^{-t(\mathop{}\!\mathbin{\bigtriangleup}+c)}) is just ectTr(et)e^{-ct}\operatorname{{Tr}}(e^{-t\mathop{}\!\mathbin{\bigtriangleup}}). If there exists a parallel section ϕ\phi for \nabla (equivalently for the connection on the dual bundle), then with Hϕ={sϕ:sΓ(E)}¯H_{\phi}=\overline{\{s\otimes\phi:s\in\Gamma(E)\}}, and appropriate normalization on ϕ\phi,

Tr(et)=Tr|Hϕ(etEnd)\displaystyle\operatorname{{Tr}}(e^{-t\mathop{}\!\mathbin{\bigtriangleup}})=\operatorname{{Tr}}|_{H_{\phi}}(e^{-t\mathop{}\!\mathbin{\bigtriangleup}^{\operatorname{{End}}}}) (4)

The parallel section, therefore, allows one to extract the heat kernel trace Tr(et)\operatorname{{Tr}}(e^{-t\mathop{}\!\mathbin{\bigtriangleup}}) from the Tr(etEnd)\operatorname{{Tr}}(e^{-t\mathop{}\!\mathbin{\bigtriangleup}^{\operatorname{{End}}}}).

In the setting of the canonical spectral triple, we specialize to the spinor bundle. The existence of a parallel spinor constrains the holonomy of Levi-Civita connection of the manifold. In particular, this implies for Riemannian (M,g)(M,g) that the Ricci tensor vanishes[8, § 6.3], therefore, D2=D^{2}=\mathop{}\!\mathbin{\bigtriangleup}.

Remark 1.3.

A remark on the Lorentzian and Kahler analogs: the existence of a parallel spinor constrains the holonomy[8], so is a strict condition. But the analysis is expected to work for Dirac operators coming from Spin\text{Spin}^{\mathbb{C}}. The extension to the complex setting, especially Kahler manifolds, should also follow easily, and these provide an interesting class of examples. In the Lorentzian setting, there’s a richer supply of parallel spinors, however, the essential difficulty there is that the spectrum of the Dirac operator is no longer discrete and the regularity requirements become unclear. The spectral action principle for Lorentzian scattering space obtained by [6] suggests the obvious question of a probabilistic interpretation in the Lorentzian setting as well. Since 𝒞(M){\mathcal{C}}(M) is a commutative algebra, so positivity is equivalent to complete positivity, therefore, by same ideas, etEnd𝒞(M)e^{-t\mathop{}\!\mathbin{\bigtriangleup}^{\operatorname{{End}}}}\subset{\mathcal{C}}(M) is a completely positive semigroup, and the question is reasonable.

Remark 1.4.

A remark on in absence of a parallel spinor: The requirement of the parallel section can be worked around sometimes, for example, S1×S3S^{1}\times S^{3} where S1S^{1} carries the flat connection. By [9], there is a quantum stochastic flow for the Laplacian for the endomorphism connection associated to the homogeneous HH-connection on the spinor bundle over S3S^{3}. Since S3S^{3} is also symmetric the homogeneous connection and the Levi-Civita connection agree. However, there are no parallel spinors on S3S^{3}, but there do exist Killing spinors with Killing constant ±1/2\pm 1/2. In such a setting, it’s possible to use the homogeneous space construction for quantum diffusion on the endomorphism bundle to get at the spectral action and the heat kernel, the idea being to modify the connection to make a Killing spinor parallel. Let ±\mathop{}\!\mathbin{\bigtriangleup}^{\prime}_{\pm} be the connection Laplacians, with k=±1/2k=\pm 1/2, for the connections ±\nabla^{\prime}_{\pm}, acting by Xϕ=Xϕ±kXϕ\nabla^{\prime}_{X}\phi=\nabla_{X}\phi\pm kX\cdot\phi where \cdot is the Clifford action, and \nabla the Levi-Civita connection, then on SnS^{n}, the Wietzenbock identity (D±k)2=±+14(n1)2(D\pm k)^{2}=\mathop{}\!\mathbin{\bigtriangleup}^{\prime}_{\pm}+\frac{1}{4}(n-1)^{2} holds (this is a calculation, that works more generally than SnS^{n}). Using that Killing spinors are parallel for ±\mathop{}\!\mathbin{\bigtriangleup}_{\pm}, Tr(et(D±k)2)=et/4(n1)2Tr(et±)\operatorname{{Tr}}(e^{-t(D\pm k)^{2}})=e^{-t/4(n-1)^{2}}\operatorname{{Tr}}(e^{-t\mathop{}\!\mathbin{\bigtriangleup}_{\pm}}) can be computed. It just needs to be checked that the flow exists for ±\mathop{}\!\mathbin{\bigtriangleup}_{\pm}. For S1×S3S^{1}\times S^{3} examples, the Dirac operator and its square can be explicitly computed by specializing the Dirac operator for Robertson-Walker metrics to a constant warping function[4].

2 The structure maps for the Laplacian generated flow

From [9], the heat semigroup etEnde^{-t\mathop{}\!\mathbin{\bigtriangleup}^{\operatorname{{End}}}} is a quantum dynamical semigroup on End(E)EE\operatorname{{End}}(E)\equiv E\otimes E^{*} with etEnd(1)=1e^{-t\mathop{}\!\mathbin{\bigtriangleup}^{\operatorname{{End}}}}({\textbf{1}})={\textbf{1}} for all tt. We will work with the semigroup living on 𝒞(M)End(E){\mathcal{C}}(M)\subset\operatorname{{End}}(E). To derive the qsde associated to the heat semigroup, we start by computing the flow the structure matrix for the associated Evans-Hudson flow following the standard prescription (see [14]): first we compute the kernel for the generator =End{\mathcal{L}}={\mathop{}\!\mathbin{\bigtriangleup}^{\operatorname{{End}}}} on the 𝒜=𝒞(M){\mathcal{A}}_{\infty}={\mathcal{C}}^{\infty}(M) acting on End(E)\operatorname{{End}}(E) defined by K:X×X(EE)K_{\mathcal{L}}:X\times X\to{\mathcal{B}}(E\otimes E^{*}) for X:=𝒜×𝒜X:={\mathcal{A}}_{\infty}\times{\mathcal{A}}_{\infty}, where KK_{\mathcal{L}} for any given any :XX{\mathcal{L}}:X\to X is defined by

X×X((f1,f2),(g1,g2))(f1f2g2g1)+f1(f2g2)g1(f1f2g2)g1f1(f2g2g1)𝒞(M)X\times X\ni((f_{1},f_{2}),(g_{1},g_{2}))\to{\mathcal{L}}(f_{1}^{*}f_{2}^{*}g_{2}g_{1})+f_{1}^{*}{\mathcal{L}}(f_{2}^{*}g_{2})g_{1}-{\mathcal{L}}(f_{1}^{*}f_{2}^{*}g_{2})g_{1}-f_{1}^{*}{\mathcal{L}}(f_{2}^{*}g_{2}g_{1})\in{\mathcal{C}}(M)

For a basis of eigensections (hi)(h_{i}) of \mathop{}\!\mathbin{\bigtriangleup}, write ihihi=1End(E)\sum_{i}h_{i}\otimes h_{i}^{*}=1\in\operatorname{{End}}(E). As Xϕ=df(X)ϕ+fX(ϕ)\nabla_{X}\phi=df(X)\phi+f\nabla_{X}(\phi), for fm,gm𝒞(M)f_{m},g_{m}\in{\mathcal{C}}(M), the contribution of ii\nabla_{i}\otimes\nabla_{i} term to the kernel ((f1,f2),(g1,g2))(f1f2g2g1)+f1(f2g2)g1(f1f2g2)g1f1(f2g2g1)((f_{1},f_{2}),(g_{1},g_{2}))\to{\mathcal{L}}(f_{1}^{*}f_{2}^{*}g_{2}g_{1})+f_{1}^{*}{\mathcal{L}}(f_{2}^{*}g_{2})g_{1}-{\mathcal{L}}(f_{1}^{*}f_{2}^{*}g_{2})g_{1}-f_{1}^{*}{\mathcal{L}}(f_{2}^{*}g_{2}g_{1}) vanishes because fm,gnf_{m},g_{n} are acting by multiplication on the Xϕ\nabla_{X}\phi term and they all commute, while for the term df(X)ϕdf(X)\phi, d(f1f2g2g1)+f1d(f2g2)g1f1d(f2g2g1)d(f1f2g2)g1d(f_{1}^{*}f_{2}^{*}g_{2}g_{1})+f_{1}^{*}d(f_{2}^{*}g_{2})g_{1}-f_{1}^{*}d(f_{2}^{*}g_{2}g_{1})-d(f_{1}^{*}f_{2}^{*}g_{2})g_{1} vanishes as well since

(f1d(f2g1g2)\displaystyle(f_{1}^{*}d(f_{2}^{*}g_{1}g_{2}) +(df1)f2g1g2)+f1d(f2g1)g2f1d(f2g2g1)d(f1f2g2)g1\displaystyle+(df_{1}^{*})f_{2}^{*}g_{1}g_{2})+f_{1}^{*}d(f_{2}^{*}g_{1})g_{2}-f_{1}^{*}d(f_{2}^{*}g_{2}g_{1})-d(f_{1}^{*}f_{2}^{*}g_{2})g_{1} (5)
=(df1)f2g1g2+f1d(f2g1)g2((df1)(f2g2)g1+f1d(f2g2)g1)=0\displaystyle=(df_{1}^{*})f_{2}^{*}g_{1}g_{2}+f_{1}^{*}d(f_{2}^{*}g_{1})g_{2}-((df_{1}^{*})(f_{2}^{*}g_{2})g_{1}+f_{1}^{*}d(f_{2}^{*}g_{2})g_{1})=0

By commutativity of 𝒞(M){\mathcal{C}}(M), 11\otimes\mathop{}\!\mathbin{\bigtriangleup} also has no contribution, since by convention the f1End=i(fhi)hif\cdot 1_{\operatorname{{End}}}=\sum_{i}(f\cdot h_{i})\otimes h_{i}^{*}. The only contribution to the kernel comes from the piece 1\mathop{}\!\mathbin{\bigtriangleup}\otimes 1. Suppose ϕ\phi is an eigensection with eigenvalue λ\lambda. Then with f𝒞(M)f\in{\mathcal{C}}^{\infty}(M), we have

fϕ\displaystyle\mathop{}\!\mathbin{\bigtriangleup}f\phi =kekekfϕ=kek(df(ek)ϕ+fekϕ)\displaystyle=-\sum_{k}\nabla_{e_{k}}\nabla_{e_{k}}f\phi=-\sum_{k}\nabla_{e_{k}}\left(df(e_{k})\phi+f\nabla_{e_{k}}\phi\right)
=k(d2f(ek,ek)ϕ+2df(ek)ekϕ+ekekϕ)=k(d2f(ek,ek)Φ+2df(ek)ekΦ)+fλϕ\displaystyle=-\sum_{k}\left(d^{2}f(e_{k},e_{k})\phi+2df(e_{k})\nabla_{e_{k}}\phi+\nabla_{e_{k}}\nabla_{e_{k}}\phi\right)=-\sum_{k}\left(d^{2}f(e_{k},e_{k})\Phi+2df(e_{k})\nabla_{e_{k}}\Phi\right)+f\lambda\phi

Again by commutativity of 𝒞(M),{\mathcal{C}}(M), λfϕ\lambda f\phi term does not contribute to the kernel, while a the same calculation as equation 5 establishes that k2df(ek)ekϕ\sum_{k}2df(e_{k})\nabla_{e_{k}}\phi contributes zero. The only contribution to the kernel comes from the term kd2f(ek,ek)Φ-\sum_{k}d^{2}f(e_{k},e_{k})\Phi. Computing the kernel for kd2f(ek,ek)\sum_{k}d^{2}f(e_{k},e_{k}) gives 2(kda1(ek)db1(ek))a2b2-2\left(\sum_{k}da_{1}(e_{k})db_{1}(e_{k})\right)a_{2}b_{2}.

2.1 The Kolmogorov decomposition

Following Goswami-Sinha construction of the flow generator, we compute the Kolmogorov decomposition for the kernel for =End/2{\mathcal{L}}=-\mathop{}\!\mathbin{\bigtriangleup}^{\operatorname{{End}}{}}/2 (to avoid factors of 2-2), the data we need will be derived from the structure theory of this kernel,

K((a1,a2),(b1,b2))=(kda1(ek)db1(ek))a2b2\displaystyle K_{{\mathcal{L}}}((a_{1},a_{2}),(b_{1},b_{2}))=\left(\sum_{k}da_{1}(e_{k})db_{1}(e_{k})\right)a_{2}b_{2} (6)

As in [14, Thm 2.2.7], the Kolmogorov decomposition is taken to be the reproducing kernel Hilbert space

R=LinSpan{K(,b)u:bC(M),uE}¯R_{{\mathcal{L}}}=\overline{\operatorname{{LinSpan}}\{K_{{\mathcal{L}}}(\cdot,b)u:b\in C^{\infty}(M),u\in E\}}

with map V:𝒞(M)(E,R)V:{\mathcal{C}}^{\infty}(M)\to{\mathcal{B}}(\mathrm{E},R_{\mathcal{L}}) for x𝒞(M)x\in{\mathcal{C}}^{\infty}(M) given by

V(x):ER,EuK(,x)uRV(x):E\to R_{\mathcal{L}},E\ni u\to K_{\mathcal{L}}(\cdot,x)u\in R_{\mathcal{L}}

By definition, K(,x)uK_{\mathcal{L}}(\cdot,x)u is total in RR_{{\mathcal{L}}} making the decomposition minimal.

Remark 2.1.

Comparing with equation 6 the term a2b2kda1(ek)db1(ek)a_{2}b_{2}\sum_{k}da_{1}(e_{k})db_{1}(e_{k}) in local coordinates can be interpret as the form a2kda1(ek)deka_{2}\sum_{k}da_{1}(e_{k})de_{k} evaluated on the vector field b2kdb1(ek)ekb_{2}\sum_{k}db_{1}(e_{k})\frac{\partial}{\partial e_{k}}. So each KL(,b)uK_{L}(\cdot,b)u is valued in Γ(TM)L2(M,E)\Gamma(TM)\otimes L^{2}(M,E) and for each x=(a1,a2)𝒞(M)×𝒞(M)x=(a_{1},a_{2})\in{\mathcal{C}}^{\infty}(M)\times{\mathcal{C}}^{\infty}(M), V(x)V(x) is a 11-form acts by contracting with Γ(TM)\Gamma(TM) component of K(,b)uK_{{\mathcal{L}}}(\cdot,b)u.

2.2 The structure matrix

The structure matrix is the map Θ:𝒜0(H(k0))\Theta:{\mathcal{A}}_{0}\to{\mathcal{B}}(H\otimes(\mathbb{C}\oplus k_{0})) for a dense subalgebra 𝒜0𝒜{\mathcal{A}}_{0}\subset{\mathcal{A}}, and Hilbert spaces H,k0H,k_{0} (alternatively, a map Θ:𝒜0𝒜0(k0),𝒜0(H)\Theta:{\mathcal{A}}_{0}\to{\mathcal{A}}_{0}\otimes{\mathcal{B}}(\mathbb{C}\oplus k_{0}),{\mathcal{A}}_{0}\subset{\mathcal{B}}(H)) given by

Θ(x)=((x)δ(x)δ(x)σ(x))\displaystyle\Theta(x)=\begin{pmatrix}{\mathcal{L}}(x)&\delta^{\dagger}(x)\\ \delta(x)&\sigma(x)\end{pmatrix} (7)

where :𝒜0𝒜0{\mathcal{L}}:{\mathcal{A}}_{0}\to{\mathcal{A}}_{0} is a *-linear map such that (xy)(x)yx(y)=δ(x)δ(y){\mathcal{L}}(xy)-{\mathcal{L}}(x)y-x{\mathcal{L}}(y)=\delta^{\dagger}(x)\delta(y), δ:𝒜0𝒜0k0\delta:{\mathcal{A}}_{0}\to{\mathcal{A}}_{0}\otimes k_{0} is a π\pi-derivation, Θ(1)=0\Theta(1)=0. Additionally, σ(x)=π(x)x1\sigma(x)=\pi(x)-x\otimes 1. The maps ,δ,σ\mathop{}\!\mathbin{\bigtriangleup},\delta,\sigma satisfy these conditions iff the first order Ito product formula holds[1, lemma 2.2].

For the structure matrix {\mathcal{L}} is the semigroup generator End-\mathop{}\!\mathbin{\bigtriangleup}^{\operatorname{{End}}}. Following [14], δ,σ\delta,\sigma are extracted from the minimal Kolmogorov decomposition: the decomposition (R,V)(R_{{\mathcal{L}}},V), induces the following maps on 𝒜:=𝒞(M)𝒞(M)=𝒜{\mathcal{A}}_{\infty}:={\mathcal{C}}^{\infty}(M)\subset{\mathcal{C}}(M)={\mathcal{A}},

ρ:𝒜(R),ρ(x)(V(,b)u)=V(,xb)u\displaystyle\rho:{\mathcal{A}}_{\infty}\to{\mathcal{B}}(R_{\mathcal{L}}),\rho(x)(V(\cdot,b)u)=V(\cdot,xb)u
α:𝒜(E,R),α(x)=V(x,1)\displaystyle\alpha:{\mathcal{A}}_{\infty}\to{\mathcal{B}}(E,R_{{\mathcal{L}}}),\alpha(x)=V(x,1)

With remark 2.1 and equation 6 in mind, ρ(f)\rho(f), f𝒞(M)f\in{\mathcal{C}}^{\infty}(M), is multiplication by ff on RR_{{\mathcal{L}}} while α(f)\alpha(f) acts by contraction with 11-form kdf(ek)dek\sum_{k}df(e_{k})de_{k}. The representation ρ\rho is the identity map: C(M)C^{\infty}(M) is interpreted as acting by multiplication on RR_{{\mathcal{L}}}, and α\alpha is a derivation.

The construction of the structure maps proceeds as in [14, Thm 6.6.1]. To start define the Hilbert 𝒜{\mathcal{A}}-module E={α(x)y:x,y𝒜}¯E=\overline{\{\alpha(x)y:x,y\in{\mathcal{A}}_{\infty}\}} where the closure is with respect to operator norm for (HE,R){\mathcal{B}}({\textsf{H}}_{E},R_{{\mathcal{L}}}). 𝒜{\mathcal{A}} has right action on EE by multiplication (where the norm density of 𝒜𝒜{\mathcal{A}}_{\infty}\subset{\mathcal{A}} is utilized) and the 𝒜{\mathcal{A}}-valued inner product is E×E(a,b)a,b=abE\times E\ni(a,b)\to{\langle a,b\rangle}=a^{*}b.

Now EE can be identified α(f)gE\alpha(f)g\in E can be identified with gdf𝒜C*W𝒜C*L2(Ω1(M))g\otimes df\in{\mathcal{A}}\otimes_{\emph{C}\textsuperscript{*}}W\subset{\mathcal{A}}\otimes_{\emph{C}\textsuperscript{*}}L^{2}(\Omega^{1}(M)) where W={df:f𝒞(M)}¯W=\overline{\{df:f\in{\mathcal{C}}^{\infty}(M)\}} (note MM is compact) is a closed subspace of L2(Ω1(M))=WWL^{2}(\Omega^{1}(M))=W\oplus W^{\perp} (the 𝒜{\mathcal{A}} valued innerproduct is pointwise tensor contraction). However, 𝒜C*L2(Ω1(M))=L2(Ω1(M)){\mathcal{A}}\otimes_{\emph{C}\textsuperscript{*}}L^{2}(\Omega^{1}(M))=L^{2}(\Omega^{1}(M)) since 𝒜=𝒞(M){\mathcal{A}}={\mathcal{C}}(M) the tensor product is 𝒞(M){\mathcal{C}}(M)-balanced: aω𝒜C*L2(Ω1(M))a\otimes\omega\in{\mathcal{A}}\otimes_{\emph{C}}\textsuperscript{*}L^{2}(\Omega^{1}(M)) is simply 1aω1\otimes a\omega.

This choice of the Hilbert space L2(Ω1(M))L^{2}(\Omega^{1}(M)) simplifies the application of Kasparov stabilization theorem[14, Thm 4.1.10]. Now the Hilbert-C*-module 𝒜C*L2(Ω1(M))=𝒜L2(Ω1(M))=L2(Ω1(M)){\mathcal{A}}\otimes_{\emph{C}\textsuperscript{*}}L^{2}(\Omega^{1}(M))={\mathcal{A}}\otimes L^{2}(\Omega^{1}(M))=L^{2}(\Omega^{1}(M)), and by Kasparov stabilization theorem, there’s a unitary map t:E𝒜L2(Ω1(M))𝒜L2(Ω1(M))t^{\prime}:E\oplus{\mathcal{A}}\otimes L^{2}(\Omega^{1}(M))\to{\mathcal{A}}\otimes L^{2}(\Omega^{1}(M)). In this case it can be explicitly computed, but it turns out not to matter.

The tt^{\prime} yields a unitary embedding t:E𝒜C*L2(Ω1(M)),t=t|Et:E\to{\mathcal{A}}\otimes_{\emph{C}\textsuperscript{*}}L^{2}(\Omega^{1}(M)),t=t^{\prime}|_{E}. Note that since the C*-module 𝒜C*L2(Ω1(M)){\mathcal{A}}\otimes_{\emph{C}\textsuperscript{*}}L^{2}(\Omega^{1}(M)) has a basis, one does not need Kasparov’s stabilization theorem and an abstract embedding, everything can be done explicitly.

Define δ(x)=t(α(x))\delta(x)=t(\alpha(x)). Now ρ\rho induces a left action ρ^\hat{\rho} on EE, ρ^(x)(α(y))=(α(xy)α(x)y)\hat{\rho}(x)(\alpha(y))=(\alpha(xy)-\alpha(x)y). But as α\alpha is a ρ\rho-derivation, ρ^(x)=xα(y)\hat{\rho}(x)=x\alpha(y), so ρ^(x)\hat{\rho}(x) is multiplication by xx and is again identity representation of 𝒜{\mathcal{A}} acting by multiplication on sections on the bundle. Set π(x)=tρ^(x)t\pi(x)=t\hat{\rho}(x)t^{*}, again π=1\pi={\textbf{1}} (so the explicit form of tt does not come into play). Therefore, with equation 7 in reference,

  • The multiplicity space k0=L2(Ω1(M))k_{0}=L^{2}(\Omega^{1}(M))

  • δ:𝒜𝒜k0,δ(f)=1df\delta:{\mathcal{A}}\to{\mathcal{A}}\otimes k_{0},\delta(f)=1\otimes df.

  • The requirement that δ(x)δ(y)=(xy)x(y)(x)y\delta(x)^{*}\delta(y)={\mathcal{L}}(x^{*}y)-x^{*}{\mathcal{L}}(y)-{\mathcal{L}}(x)^{*}y forces δ(f):=δ(f)\delta^{\dagger}(f):=\delta(f^{*})^{*} to act by δ(f)(1ω)=df,ω=idf(ei)ω(ei)𝒞(M)\delta(f)^{\dagger}(1\otimes\omega)={\langle df,\omega\rangle}=\sum_{i}df(e_{i})\omega(e_{i})\in{\mathcal{C}}(M) (using f=ff=f^{*}); so δ(f)(aω)=adf,ω\delta(f)^{\dagger}(a\otimes\omega)=a{\langle df,\omega\rangle} works.

  • σ=0\sigma=0 since σ(x)=π(x)x1k0\sigma(x)=\pi(x)-x\otimes{\textbf{1}}_{k_{0}} since π:𝒜0𝒜0(k0)\pi:{\mathcal{A}}_{0}\to{\mathcal{A}}_{0}\otimes{\mathcal{B}}(k_{0}) is identity

  • {\mathcal{L}} is the generator End/2-\mathop{}\!\mathbin{\bigtriangleup}^{\operatorname{{End}}}/2 for the semigroup etEnd/2e^{-t\mathop{}\!\mathbin{\bigtriangleup}^{\operatorname{{End}}}/2}.

Therefore, the structure matrix for the Laplacian generated flow is the map

𝒜0fΘ(f)=(End(f1)/2δ(f)1df0)(H(k0))\displaystyle{\mathcal{A}}_{0}\ni f\to\Theta(f)=\begin{pmatrix}-\mathop{}\!\mathbin{\bigtriangleup}^{\operatorname{{End}}}(f\cdot{\textbf{1}})/2&\delta^{\dagger}(f)\\ 1\otimes df&0\end{pmatrix}\in{\mathcal{B}}(H\otimes(\mathbb{C}\oplus k_{0})) (8)

3 Growth bounds and continuity

Let (M,g)(M,g) be a Ricci-flat Riemannian manifold with Levi-Civita connection \nabla. Since action of End\mathop{}\!\mathbin{\bigtriangleup}^{\operatorname{{End}}} on 𝒞(M){\mathcal{C}}(M) can be identified with the Laplace-Beltrami operator M\mathop{}\!\mathbin{\bigtriangleup}^{M}\equiv\mathop{}\!\mathbin{\bigtriangleup} (and the lift, the rough Laplacian, to the tensor bundle), we will work with \mathop{}\!\mathbin{\bigtriangleup}. Let ϕi\phi_{i} be the an eigenfunction of the Laplace-Beltrami operator, =gabab\mathop{}\!\mathbin{\bigtriangleup}=-g^{ab}\nabla_{a}\nabla_{b}, with eigenvalue λi2\lambda_{i}^{2}, ϕi=λi2ϕi\mathop{}\!\mathbin{\bigtriangleup}\phi_{i}=\lambda_{i}^{2}\phi_{i}. Define {\mathcal{F}} as the generated by finite products and sums of ϕi\phi_{i}’s,

=LinSpan(i[k]ϕi:k){\mathcal{F}}=\operatorname{{LinSpan}}\left(\prod_{i\in[k]}\phi_{i}:k\in{\mathds{N}}\right)

{\mathcal{F}} is norm-dense inside 𝒞(M){\mathcal{C}}(M), and 1:={i[k]ϕi:k=1}{\mathcal{F}}_{1}:=\{\prod_{i\in[k]}\phi_{i}:k=1\}\subset{\mathcal{F}} forms a basis for 𝒞(M){\mathcal{C}}(M). Set d={df:f1}k0d{\mathcal{F}}=\{df:f\in{\mathcal{F}}_{1}\}\subset k_{0} as a norm-dense subspace, (k0)(k_{0})_{\infty}. The constructions from [14], [1] both proceed by controlling the growth of the flow generator on a dense algebra, a role here played by {\mathcal{F}}. Note we are only interested in bound on the \mathop{}\!\mathbin{\bigtriangleup}, we can work with either =±gabab\mathop{}\!\mathbin{\bigtriangleup}=\pm g^{ab}\nabla_{a}\nabla_{b} to avoid tracking signs if needed.

3.1 Growth of Laplacian iterates

For xMx\in M fix a small neighborhood of xx, VV, with UVU\subset V open, such that U¯V\bar{U}\subset V, with coordinates (xi)(x^{i}) and (xi)i(\partial_{x^{i}})\equiv\partial_{i} the coordinate vector fields. For any multi-index α\alpha we denote α=αkα1\partial^{\alpha}=\partial_{\alpha_{k}}\dots\partial_{\alpha_{1}}, and same for α\nabla_{\alpha} with ii\nabla_{i}\equiv\nabla_{\partial_{i}}. Ricci-flatness implies for all fC(M)f\in C^{\infty}(M) (see, for instance, [5, lemma 1.36]),

[,i](f)=Ricikk(f)=0\displaystyle[\mathop{}\!\mathbin{\bigtriangleup},\nabla_{i}](f)=\text{Ric}_{i}^{k}\nabla_{k}(f)=0 (9)

On (p,q)(p,q)-tensors s,ss,s^{\prime} there’s a natural innerproduct by contraction with gij,gij,s,s=gi1j1gm1n1si1m1sj1n1g^{ij},g_{ij},{\langle s,s^{\prime}\rangle}=g^{i_{1}j_{1}}\dots g_{m_{1}n_{1}}{s^{\prime}}^{m_{1}\dots}_{i_{1}\dots}s^{n_{1}\dots}_{j_{1}\dots}. The Levi-Civita connection has a lift to the tensor bundle and an associated connection Laplacian, both also denoted ,\nabla,\mathop{}\!\mathbin{\bigtriangleup}. Denote by ku\nabla^{k}u the kthk^{th}-covariant derivative and define the point-wise length with the innerproduct[10, § 2.2.1]:

(ku)2=gi1j1gikjk(ku)i1ik(ku)j1jk=ku,ku\displaystyle\ell(\nabla^{k}u)^{2}=g^{i_{1}j_{1}}\dots g^{i_{k}j_{k}}(\nabla^{k}u)_{i_{1}\dots i_{k}}(\nabla^{k}u)_{j_{1}\dots j_{k}}={\langle\nabla^{k}u,\nabla^{k}u\rangle} (10)
Lemma 3.1.

For all f𝒞(M)f\in{\mathcal{C}}^{\infty}(M),

[,i]jkjk1j1(f)=0[\mathop{}\!\mathbin{\bigtriangleup},\nabla_{i}]\nabla_{j_{k}}\nabla_{j_{k-1}}\dots\nabla_{j_{1}}(f)=0
Proof.

Using

(ku)i1ik=i1i2ik(u)\displaystyle(\nabla^{k}u)_{i_{1}\dots i_{k}}=\nabla_{i_{1}}\nabla_{i_{2}}\dots\nabla_{i_{k}}(u) (11)

and that gijg^{ij} commutes with ,\mathop{}\!\mathbin{\bigtriangleup},\nabla since \nabla is metric compatible,

([,i]jkjk1j1(f))2\displaystyle\ell([\mathop{}\!\mathbin{\bigtriangleup},\nabla_{i}]\nabla_{j_{k}}\nabla_{j_{k-1}}\dots\nabla_{j_{1}}(f))^{2} =giigj1j1gjkjk([,i]jkj1(f))([,i]jkj1(f))\displaystyle=g^{ii^{\prime}}g^{j_{1}j_{1}^{\prime}}\dots g^{j_{k}j_{k}^{\prime}}([\mathop{}\!\mathbin{\bigtriangleup},\nabla_{i}]\nabla_{j_{k}}\dots\nabla_{j_{1}}(f))([\mathop{}\!\mathbin{\bigtriangleup},\nabla_{i}^{\prime}]\nabla_{j^{\prime}_{k}}\dots\nabla_{j^{\prime}_{1}}(f))
=gii([,i][,i])(gj1j1gjkjkjkj1(f)jkj1(f))\displaystyle=g^{ii^{\prime}}([\mathop{}\!\mathbin{\bigtriangleup},\nabla_{i}][\mathop{}\!\mathbin{\bigtriangleup},\nabla_{i^{\prime}}])(g^{j_{1}j_{1}^{\prime}}\dots g^{j_{k}j_{k}^{\prime}}\nabla_{j_{k}}\dots\nabla_{j_{1}}(f)\nabla_{j^{\prime}_{k}}\dots\nabla_{j^{\prime}_{1}}(f)) (12)

Since h:=gj1j1gjkjkjkj1(f)jkj1(f)𝒞h:=g^{j_{1}j_{1}^{\prime}}\dots g^{j_{k}j_{k}^{\prime}}\nabla_{j_{k}}\dots\nabla_{j_{1}}(f)\nabla_{j^{\prime}_{k}}\dots\nabla_{j^{\prime}_{1}}(f)\in{\mathcal{C}}^{\infty}, from equation 9,

([,i]jkjk1j1(f))2=gii([,i]([,i])(h)=0.\ell([\mathop{}\!\mathbin{\bigtriangleup},\nabla_{i}]\nabla_{j_{k}}\nabla_{j_{k-1}}\dots\nabla_{j_{1}}(f))^{2}=g^{ii^{\prime}}([\mathop{}\!\mathbin{\bigtriangleup},\nabla_{i}]([\mathop{}\!\mathbin{\bigtriangleup},\nabla_{i^{\prime}}])(h)=0.

Remark 3.2.

Similar bounds can be obtained for Einstein manifolds where the Ricci tensor is proportional to the metric. But since the analysis here needs a parallel spinor, the Evans-Hudson qsde over Einstein manifolds are not considered.

Proposition 3.3.

For all f𝒞(M)f\in{\mathcal{C}}^{\infty}(M), (f)dimM(2(f))\left\lVert\mathop{}\!\mathbin{\bigtriangleup}(f)\right\rVert\leq\sqrt{\dim M}\ell(\nabla^{2}(f)) and k(f)(dimM)k/2(k(f))\left\lVert\mathop{}\!\mathbin{\bigtriangleup}^{k}(f)\right\rVert\leq(\dim M)^{k/2}\ell(\nabla^{k}(f))

Proof.

Using (gab)2=gacgbdgabgcd=gcdgcd=dimM\ell(g_{ab})^{2}=g^{ac}g^{bd}g_{ab}g_{cd}=g^{cd}g_{cd}=\dim M

f=gababf=gcagbdgcdabf=gcd,abf(gcd)(abf)dimM(abf)\mathop{}\!\mathbin{\bigtriangleup}f=g^{ab}\nabla_{a}\nabla_{b}f=g^{ca}g^{bd}g_{cd}\nabla_{a}\nabla_{b}f={\langle g_{cd},\nabla_{a}\nabla_{b}f\rangle}\leq\ell(g_{cd})\ell(\nabla_{a}\nabla_{b}f)\leq\sqrt{\dim M}\ell(\nabla_{a}\nabla_{b}f)

where we used Cauchy-Schwarz. The bound (dimM)k/2(k(f))(\dim M)^{k/2}\ell(\nabla^{k}(f)) follows identically using (gi1j1gikjk)2=(dimM)k\ell(g_{i_{1}j_{1}}\dots g_{i_{k}j_{k}})^{2}=(\dim M)^{k}. ∎

The growth of Laplacian and its powers is clear on ϕi\phi_{i}’s, since k(ϕi)=λi2kϕi\mathop{}\!\mathbin{\bigtriangleup}^{k}(\phi_{i})=\lambda_{i}^{2k}\phi_{i}. Controlling the growth on products of ϕi\phi_{i}’s will require control over

kϕi,ϕjk(x)kϕi(x)2kϕj(x)=(kϕi(x))(kϕj(x))\displaystyle{\langle\nabla^{k}\phi_{i},\nabla\phi^{k}_{j}\rangle}(x)\leq\sqrt{\left\lVert\nabla^{k}\phi_{i}(x)\right\rVert_{2}\left\lVert\nabla^{k}\phi_{j}(x)\right\rVert}=\ell(\nabla^{k}\phi_{i}(x))\ell(\nabla^{k}\phi_{j}(x)) (13)

(using eq 10 with k=1k=1 and Cauchy-Schwartz inequality). To see this note how Laplacian and its iterated powers act on products of ϕi\phi_{i}’s.

Proposition 3.4.

[Laplacian on products] For ϕj,ϕi\phi_{j},\phi_{i},

  1. i

    (ϕiϕj)=ϕi(ϕj)+ϕi(ϕj)+2ϕi,ϕj\mathop{}\!\mathbin{\bigtriangleup}(\phi_{i}\phi_{j})=\phi_{i}\mathop{}\!\mathbin{\bigtriangleup}(\phi_{j})+\phi_{i}\mathop{}\!\mathbin{\bigtriangleup}(\phi_{j})+2{\langle\nabla\phi_{i},\nabla\phi_{j}\rangle}

  2. ii

    For any kk,

    kϕi,kϕj=kϕi,kϕj+kϕi,kϕj+2k+1ϕi,k+1ϕj\displaystyle\mathop{}\!\mathbin{\bigtriangleup}{\langle\nabla^{k}\phi_{i},\nabla^{k}\phi_{j}\rangle}={\langle\nabla^{k}\mathop{}\!\mathbin{\bigtriangleup}\phi_{i},\nabla^{k}\phi_{j}\rangle}+{\langle\nabla^{k}\phi_{i},\nabla^{k}\mathop{}\!\mathbin{\bigtriangleup}\phi_{j}\rangle}+2{\langle\nabla^{k+1}\phi_{i},\nabla^{k+1}\phi_{j}\rangle} (14)
Proof.

These are straightforward computations, see appendix A for details. ∎

Therefore, it’s sufficient to bound k(ϕj)\left\lVert\nabla^{k}(\phi_{j})\right\rVert_{\infty}. For ϕj\phi_{j} with λj1\lambda_{j}\geq 1, the bound follows from [3, lemma 2.7]:

Theorem 3.5 ([3]).

With λ1\lambda\geq 1,

λj[λ,λ+1](kϕj(x))2CM,xλ2k+dimM1\displaystyle\sum_{\lambda_{j}\in[\lambda,\lambda+1]}\ell(\nabla^{k}\phi_{j}(x))^{2}\leq C_{M,x}\lambda^{2k+\dim M-1} (15)

The following corollary is immediate since MM is compact and CM=(λj+1)dimM1supxMCM,xC_{M}=(\lambda_{j}+1)^{\dim M-1}\sup_{x\in M}C_{M,x} and Kj=(λj+1)2K_{j}=(\lambda_{j}+1)^{2} can be used.

Corollary 3.6.

For MM compact, for any ϕj\phi_{j} with λj1\lambda_{j}\geq 1, (kϕj(x))2CMKjk\ell(\nabla^{k}\phi_{j}(x))^{2}\leq C_{M}K_{j}^{k}.

One expects that kϕjL2(M)\left\lVert\nabla^{k}\phi_{j}\right\rVert_{L^{2}(M)} should decay to zero for jj with λj2<1\lambda^{2}_{j}<1, and since ϕj\phi_{j}’s are smooth this is enough to establish a uniform bound. However, this will need to be leveraged locally and the boundary for the local chart will need to be taken into account. Recall the integration on parts formula for tensor fields[12, pg 50,149] when MM does have a boundary,

MF,G𝑑Vg=MFN,G𝑑Vg^MF,div(G)𝑑Vg\displaystyle\int_{M}{\langle\nabla F,G\rangle}dV_{g}=\int_{\partial M}{\langle F\otimes N^{\flat},G\rangle}dV_{\hat{g}}-\int_{M}{\langle F,\text{div}(G)\rangle}dV_{g} (16)

where g^\hat{g} is the induced metric on M,dVg,dVg^\partial M,dV_{g},dV_{\hat{g}} the associated volume forms, \cdot^{\flat} the musical isomorphism, NN the outward unit normal at M\partial M, and F,GF,G tensor fields, div(G)=Trg(G)\text{div}(G)=\operatorname{{Tr}}_{g}(\nabla G), the trace being over the last two indices. Note if G=HG=\nabla H then, div(G)=(H)-\text{div}(G)=\mathop{}\!\mathbin{\bigtriangleup}(H).

Proposition 3.7.

For u=ϕju=\phi_{j}, with jj such that λj2<1\lambda^{2}_{j}<1, kuk1u\left\lVert\nabla^{k}u\right\rVert_{\infty}\leq\left\lVert\nabla^{k-1}u\right\rVert_{\infty}

Proof.

Suppose for some xMx\in M, ku(x),ku(x)k1h(x),k1u(x)>0{\langle\nabla^{k}u(x),\nabla^{k}u(x)\rangle}-{\langle\nabla^{k-1}h(x),\nabla^{k-1}u(x)\rangle}>0. Then since uu is smooth, there exists an open neighborhood UU of xx such that on UU, ku,kuk1u,k1u>0{\langle\nabla^{k}u,\nabla^{k}u\rangle}-{\langle\nabla^{k-1}u,\nabla^{k-1}u\rangle}>0. Now let ψ\psi be such that supp(ψ)U\text{supp}(\psi)\subset U is compact, ψ0\psi\geq 0 on UU and ψ>0\psi>0 on open VUV\subset U, then

Mψku,kuψk1u,k1udVg\displaystyle\int_{M}\psi{\langle\nabla^{k}u,\nabla^{k}u\rangle}-\psi{\langle\nabla^{k-1}u,\nabla^{k-1}u\rangle}dV_{g} =Uψku,kuψk1u,k1udVg>0\displaystyle=\int_{U}{\langle\psi\nabla^{k}u,\nabla^{k}u\rangle}-\psi{\langle\nabla^{k-1}u,\nabla^{k-1}u\rangle}dV_{g}>0 (17)

Now ψku,ku=(ψk1u),kuψk1u,ku{\langle\psi\nabla^{k}u,\nabla^{k}u\rangle}={\langle\nabla(\psi\nabla^{k-1}u),\nabla^{k}u\rangle}-{\langle\nabla\psi\cdot\nabla^{k-1}u,\nabla^{k}u\rangle}, and for the first term

U(ψk1u),ku𝑑Vg=U,𝑑Vg^+Uψk1u,div(ku)𝑑Vg\int_{U}{\langle\nabla(\psi\nabla^{k-1}u),\nabla^{k}u\rangle}dV_{g}=\int_{\partial U}{\langle\cdot,\cdot\rangle}dV_{\hat{g}}+\int_{U}{\langle\psi\nabla^{k-1}u,-\text{div}(\nabla^{k}u)\rangle}dV_{g}

where U,𝑑Vg^=0\int_{\partial U}{\langle\cdot,\cdot\rangle}dV_{\hat{g}}=0 since ψ=0\psi=0 on U\partial U, while div(ku)=k1u=k1u-\text{div}(\nabla^{k}u)=\mathop{}\!\mathbin{\bigtriangleup}\nabla^{k-1}u=\nabla^{k-1}\mathop{}\!\mathbin{\bigtriangleup}u using MM is Ricci flat. Therefore, we have

Mψku,ku\displaystyle\int_{M}\psi{\langle\nabla^{k}u,\nabla^{k}u\rangle} =Uψk1u,k1uUψk1u,ku𝑑Vg\displaystyle=\int_{U}\psi{\langle\nabla^{k-1}u,\nabla^{k-1}\mathop{}\!\mathbin{\bigtriangleup}u\rangle}-\int_{U}{\langle\nabla\psi\cdot\nabla^{k-1}u,\nabla^{k}u\rangle}dV_{g}
=λ2Uψk1u,k1usupp(ψ)ψk1u,ku𝑑Vg\displaystyle=\lambda^{2}\int_{U}\psi{\langle\nabla^{k-1}u,\nabla^{k-1}u\rangle}-\int_{\text{supp}(\nabla\psi)}{\langle\nabla\psi\cdot\nabla^{k-1}u,\nabla^{k}u\rangle}dV_{g}

This yields

0<Uψ\displaystyle 0<\int_{U}\psi ku,kuψk1u,k1udVg\displaystyle{\langle\nabla^{k}u,\nabla^{k}u\rangle}-\psi{\langle\nabla^{k-1}u,\nabla^{k-1}u\rangle}dV_{g} (18)
=λ2Uψk1u,k1usupp(ψ)ψk1u,ku𝑑VgUψk1u,k1u𝑑Vg\displaystyle=\lambda^{2}\int_{U}\psi{\langle\nabla^{k-1}u,\nabla^{k-1}u\rangle}-\int_{\text{supp}(\nabla\psi)}{\langle\nabla\psi\cdot\nabla^{k-1}u,\nabla^{k}u\rangle}dV_{g}-\int_{U}\psi{\langle\nabla^{k-1}u,\nabla^{k-1}u\rangle}dV_{g}
=Uψ(λ21)k1u2𝑑Vgsupp(ψ)ψk1u,ku𝑑Vg\displaystyle=\int_{U}\psi(\lambda^{2}-1)\left\lVert\nabla^{k-1}u\right\rVert^{2}dV_{g}-\int_{\text{supp}(\nabla\psi)}{\langle\nabla\psi\cdot\nabla^{k-1}u,\nabla^{k}u\rangle}dV_{g} (19)

Define the linear functional ω(ψ):=supp(ψ)ψk1u,ku𝑑Vg\omega(\psi):=\int_{\text{supp}(\nabla\psi)}{\langle\nabla\psi\cdot\nabla^{k-1}u,\nabla^{k}u\rangle}dV_{g}. Note

ψk1u,ku\displaystyle{\langle\nabla\psi\cdot\nabla^{k-1}u,\nabla^{k}u\rangle} =gi1j1i1ψ(gi2j2gikjk(k1u)i2ik(ku)j1j2jk)\displaystyle=g^{i_{1}j_{1}}\nabla_{i_{1}}\psi\left(g^{i_{2}j_{2}}\dots g^{i_{k}j_{k}}(\nabla^{k-1}u)_{i_{2}\dots i_{k}}(\nabla^{k}u)_{j_{1}j_{2}\dots j_{k}}\right)
=gi1j1i1ψ(gi2j2gikjk(i2iku)(j1j2jku))\displaystyle=g^{i_{1}j_{1}}\nabla_{i_{1}}\psi\left(g^{i_{2}j_{2}}\dots g^{i_{k}j_{k}}(\nabla_{i_{2}}\dots\nabla_{i_{k}}u)(\nabla_{j_{1}}\nabla_{j_{2}}\dots\nabla_{j_{k}}u)\right)
=12gi1j1i1ψj1((k1u,k1u)2)=12ψ,(k1u,k1u)2\displaystyle=\textstyle{\frac{1}{2}}g^{i_{1}j_{1}}\nabla_{i_{1}}\psi\nabla_{j_{1}}(\ell(\nabla^{k-1}u,\nabla^{k-1}u)^{2})=\textstyle{\frac{1}{2}}{\langle\nabla\psi,\nabla\ell(\nabla^{k-1}u,\nabla^{k-1}u)^{2}\rangle} (20)

By showing that there exists a ψ\psi that makes ω(ψ)0\omega(\psi)\geq 0, since (λ21)<0(\lambda^{2}-1)<0, from equation 19 it will follow that equation 18 cannot hold.

Assume that UU is small enough to be covered by a geodesic normal coordinates, and consider polar coordinates on UU centered at xx. Define τsc\tau^{c}_{s} on UU for c,s>0c,s\in\mathds{R}_{>0} such that τs(x)=c\tau_{s}(x)=c and then decays linearly in radially outwards direction with slope s-s to 0 at BR(x)\partial B_{R}(x) with c,sc,s such that supp(τsc)¯U\overline{\text{supp}(\tau^{c}_{s})}\subset U, RR depending on c,sc,s. Then τsc\tau^{c}_{s} is continuous, piecewise continuously differentiable, with compact support in UU, so weakly-differentiable, and τsc=s1R(x)\nabla\tau^{c}_{s}=-s{\textbf{1}}_{{\mathcal{B}}_{R}(x)} (there’s enough slack to work with mollified versions of τ\tau’s, but weak-differentiablility suffices for simplicity). If for some τsc\tau^{c}_{s}, ω(τsc)0\omega(\tau^{c}_{s})\geq 0 then that ψ=τsc\psi=\tau^{c}_{s} is the required ψ\psi.

If not, then ω(τsc)<0\omega(\tau^{c}_{s})<0 for all cc small enough to have support in UU. By rescaling wlog assume c=s=1c=s=1, and set τ1:=τ11\tau_{1}:=\tau^{1}_{1} ( otherwise the constants are messy). For such τ1\tau_{1}, define τ1\tau^{\prime}_{1} such that τ1(x)=0,\tau^{\prime}_{1}(x)=0, and τ1\tau^{\prime}_{1} increases linearly to 11 at B1(x)\partial B_{1}(x), and outside of B1(x)¯,τ1=0\overline{B_{1}(x)},\tau^{\prime}_{1}=0. Then ω(τ1)=ω(τ1)=δ>0\omega(\tau^{\prime}_{1})=-\omega(\tau_{1})=\delta>0 since τ1=τ1\nabla\tau_{1}^{\prime}=-\nabla\tau_{1} on supp(τ1)=supp(τ1)\text{supp}(\nabla\tau_{1}^{\prime})=\text{supp}(\nabla\tau_{1}). It remains to make τ1\tau_{1}^{\prime} continuous without changing ω(τ1)\omega(\tau^{\prime}_{1}) too much. For this set τ1,r′′=τ1\tau_{1,r}^{\prime\prime}=\tau_{1}^{\prime} on B1(x)B_{1}(x), τ1,r′′=0\tau_{1,r}^{\prime\prime}=0 on B1+r(x)cB_{1+r}(x)^{c}, and on B1+r(x)cB1(x)B_{1+r}(x)^{c}\setminus B_{1}(x), τ1,r′′\tau^{\prime\prime}_{1,r} decays linearly to 0 on B1+r(x)\partial B_{1+r}(x). Finally, since for all r>0r>0 small enough, τ1,r′′\tau^{\prime\prime}_{1,r} is piecewise continuous, continuously differentiable and compactly supported in UU, it remains to check for any ϵ>0\epsilon>0 there exists rϵ>0r_{\epsilon}>0 such that for all r<rϵr<r_{\epsilon}, ω(τs)ω(τs,r′′)ϵ\left\lVert\omega(\tau^{\prime}_{s})-\omega(\tau^{\prime\prime}_{s,r})\right\rVert\leq\epsilon. Note that

2ω(τ1)ω(τ1,r′′)=B1+r(x)(x)B1(x)τs′′,div(((k1(u))2))𝑑Vg\displaystyle 2\left\lVert\omega(\tau^{\prime}_{1})-\omega(\tau^{\prime\prime}_{1,r})\right\rVert=\left\lVert\int_{B_{1+r}(x)(x)\setminus B_{1}(x)}{\langle\tau_{s}^{\prime\prime},-\text{div}(\nabla(\ell(\nabla^{k-1}(u))^{2}))\rangle}dV_{g}\right\rVert (21)

using equation 20 and so ω(τ1)ω(τ1,r′′)ϵ\left\lVert\omega(\tau^{\prime}_{1})-\omega(\tau^{\prime\prime}_{1,r})\right\rVert\leq\epsilon for small rr because div(((k1(u))2))=((k1(u))2),-\text{div}(\nabla(\ell(\nabla^{k-1}(u))^{2}))=\mathop{}\!\mathbin{\bigtriangleup}(\ell(\nabla^{k-1}(u))^{2}), τs,r′′\tau^{\prime\prime}_{s,r} being continuous (with MM compact) are bounded. ∎

The idea above generalizes to all ϕi\phi_{i}’s giving a simple proof for growth bounds on covariant derivatives from [3] in the Ricci flat setting.

Corollary 3.8.

For u=ϕju=\phi_{j}, with λ2:=λj2\lambda^{2}:=\lambda^{2}_{j}, kuλ2k1u\left\lVert\nabla^{k}u\right\rVert_{\infty}\leq\lambda^{2}\left\lVert\nabla^{k-1}u\right\rVert_{\infty}

Proof.

Assume not, then on some open UMU\subset M, for all xUx\in U, for some c>1c>1,

ku(x),ku(x)cλ2k1h(x),k1u(x)>0{\langle\nabla^{k}u(x),\nabla^{k}u(x)\rangle}-c\lambda^{2}{\langle\nabla^{k-1}h(x),\nabla^{k-1}u(x)\rangle}>0

and as in proposition 3.7 for some ψ0\psi\geq 0 compactly supported in U,ψ>0U,\psi>0 on an open set,

Uψku(x),ku(x)𝑑VgUcλ2ψk1h(x),k1u(x)𝑑Vg>0\displaystyle\int_{U}{\langle\psi\nabla^{k}u(x),\nabla^{k}u(x)\rangle}dV_{g}-\int_{U}c\lambda^{2}\psi{\langle\nabla^{k-1}h(x),\nabla^{k-1}u(x)\rangle}dV_{g}>0 (22)
with Uψku(x),ku(x)𝑑Vg=U(ψk1u(x)),ku(x)𝑑VgUψk1u,ku𝑑Vg\text{with }\int_{U}{\langle\psi\nabla^{k}u(x),\nabla^{k}u(x)\rangle}dV_{g}=\int_{U}{\langle\nabla(\psi\nabla^{k-1}u(x)),\nabla^{k}u(x)\rangle}dV_{g}-\int_{U}{\langle\nabla\psi\cdot\nabla^{k-1}u,\nabla^{k}u\rangle}dV_{g}

Therefore, equation 22 becomes

Uλ2\displaystyle\int_{U}\lambda^{2} ψk1u(x),k1u(x)dVgUψk1u,ku𝑑VgUcλ2ψk1h(x),k1u(x)𝑑Vg\displaystyle{\langle\psi\nabla^{k-1}u(x),\nabla^{k-1}u(x)\rangle}dV_{g}-\int_{U}{\langle\nabla\psi\cdot\nabla^{k-1}u,\nabla^{k}u\rangle}dV_{g}-\int_{U}c\lambda^{2}\psi{\langle\nabla^{k-1}h(x),\nabla^{k-1}u(x)\rangle}dV_{g}
=λ2(1c)Uψk1u(x),k1u(x)𝑑VgUψk1u,ku𝑑Vg>0\displaystyle=\lambda^{2}(1-c)\int_{U}{\langle\psi\nabla^{k-1}u(x),\nabla^{k-1}u(x)\rangle}dV_{g}-\int_{U}{\langle\nabla\psi\cdot\nabla^{k-1}u,\nabla^{k}u\rangle}dV_{g}>0

But choosing ψ\psi as in 3.7, since λ2(1c)<0\lambda^{2}(1-c)<0, the last inequality cannot hold. ∎

Therefore, kϕi\left\lVert\nabla^{k}\phi_{i}\right\rVert_{\infty} is uniformly bounded for all ϕi\phi_{i}. Collecting this with propositions 3.3 and 3.4 along with corollary 3.6, yields easily implies the following bounds.

Theorem 3.9.

For all k,ϕk\in{\mathds{N}},\phi\in{\mathcal{F}},

  1. i

    There exists constants Cϕ,KϕC_{\phi},K_{\phi} such that k(ϕ)CϕKϕk\left\lVert\mathop{}\!\mathbin{\bigtriangleup}^{k}(\phi)\right\rVert\leq C_{\phi}K_{\phi}^{k}

  2. ii

    For any k,Nk,N\in{\mathds{N}}, the map k1MatN:MatN()MatN(𝒞(M))\mathop{}\!\mathbin{\bigtriangleup}^{k}\otimes{\textbf{1}}_{\textsc{Mat}_{N}}:\textsc{Mat}_{N}({\mathcal{F}})\to\textsc{Mat}_{N}({\mathcal{C}}^{\infty}(M)), k1MatN(ξ)CξKξk\left\lVert\mathop{}\!\mathbin{\bigtriangleup}^{k}\otimes{\textbf{1}}_{\textsc{Mat}_{N}}(\xi)\right\rVert\leq C_{\xi}K^{k}_{\xi} for ξ=[ξij]MatN()\xi=[\xi_{ij}]\in\textsc{Mat}_{N}({\mathcal{F}}). The same holds for (1MatN)k(\mathop{}\!\mathbin{\bigtriangleup}\otimes{\textbf{1}}_{\textsc{Mat}_{N}})^{k}.

Proof.

Part i)i) follows from estimates on k\mathop{}\!\mathbin{\bigtriangleup}^{k}, how k\mathop{}\!\mathbin{\bigtriangleup}^{k} acts on products of ϕi\phi_{i}’s and \nabla being a derivation. For part ii)ii), by growth bounds on k\mathop{}\!\mathbin{\bigtriangleup}^{k}, k(ξij)CijMijk\mathop{}\!\mathbin{\bigtriangleup}^{k}(\xi_{ij})\leq C_{ij}M_{ij}^{k} for some constants Cij,MijC_{ij},M_{ij}. so

k1MatN(ξ)=[k(ξij)]maxi(jCij(jMij)k).\displaystyle\left\lVert\mathop{}\!\mathbin{\bigtriangleup}^{k}\otimes{\textbf{1}}_{\textsc{Mat}_{N}}(\xi)\right\rVert=\left\lVert[\mathop{}\!\mathbin{\bigtriangleup}^{k}(\xi_{ij})]\right\rVert\leq\max_{i}\left(\sum_{j}C_{ij}\cdot\left(\sum_{j}M_{ij}\right)^{k}\right). (23)

Note that theorem 3.9, ii)ii) is needed to guarantee that the map-valued process JtJ_{t} evaluated on exponential vectors is pointwise bounded.

4 Picard iterates for the Laplacian generated flow

We now specialize the map-valued Evans-Hudson qsde’s to the Laplacian flow generator, Θ\Theta: from equation 8,

𝒜fΘ(f)=((f)/2δ(f)1df0)(H(k0))\displaystyle{\mathcal{A}}_{\infty}\supset{\mathcal{F}}\ni f\to\Theta(f)=\begin{pmatrix}-\mathop{}\!\mathbin{\bigtriangleup}(f)/2&\delta^{\dagger}(f)\\ 1\otimes df&0\end{pmatrix}\in{\mathcal{B}}(H\otimes(\mathbb{C}\oplus k_{0})) (24)

where k0=L2(Ω1(M))k_{0}=L^{2}(\Omega^{1}(M)), the noise space, and H=L2(S)H=L^{2}(S), the square integrable spinors on which 𝒞(M){\mathcal{C}}(M) acts. Recall (k0):=d={df:f}L2(Ω1(M)(k_{0})_{\infty}:=d{\mathcal{F}}=\{df:f\in{\mathcal{F}}\}\subset L^{2}(\Omega^{1}(M). k=L2(+,k0)k=L^{2}(\mathds{R}_{+},k_{0}) For fk0f\in\mathbb{C}\oplus k_{0}. For fk0f\in k_{0}, set f^:=1fk0\hat{f}:=1\oplus f\in\mathbb{C}\oplus k_{0} and identify ff with 0fk00\oplus f\in\mathbb{C}\oplus k_{0}

Towards establishing the existence of the quantum flow of Evans-Hudson type under the much weaker regularity assumption and without the Frechet space scaffolding from [14, Thm 8.1.38]; the usual ideas from the literature can be leveraged, however, crucial estimates need to be made from ground up which we establish next. Only a sketch of known results into which these estimates are plugged are included how they made fit into the scheme. A crucial point is an extension of the squareroot trick, which is needed since in general the algebra {\mathcal{F}}, the algebra consisting of finite products and sums of eigenfunctions of \mathop{}\!\mathbin{\bigtriangleup}, is not closed under squareroots since derivatives of squareroots grow factorially while on {\mathcal{F}} the covariant derivatives satisfy growth bounds of type (kf)CfMfk\ell(\nabla^{k}f)\leq C_{f}M_{f}^{k}. For background on map-valued qsdes, we refer to the included appendix B. To start, recall the following estimates for map-valued processes, aδ,aδ,Ia_{\delta},a_{\delta}^{\dagger},I_{\mathcal{L}} being the fundamental processes (appendix equations 40,41):

Estimate 4.1.

[14, Thm 5.4.7,8.1.37] For a map-valued integrable process YsY_{s},

0tYs(aδ+I)(ds)(xE(f))u2\displaystyle\left\lVert\int_{0}^{t}Y_{s}\circ(a_{\delta}+I_{\mathcal{L}})(ds)(x\otimes\textsc{E}(f))u\right\rVert^{2} et0tYs((x)+δ(x),f(s))E(f))u2ds\displaystyle\leq e^{t}\int_{0}^{t}\left\lVert Y_{s}({\mathcal{L}}(x)+{\langle\delta(x^{*}),f(s)\rangle})\otimes\textsc{E}(f))u\right\rVert^{2}ds (25)
0tYs(aδ)(ds)(xE(f))u2\displaystyle\left\lVert\int_{0}^{t}Y_{s}\circ(a^{\dagger}_{\delta})(ds)(x\otimes\textsc{E}(f))u\right\rVert^{2} et0t(Ys~(δ(x)E(f))u2+Ys(f(s),δ(x))E(f)u2)𝑑s\displaystyle\leq e^{t}\int_{0}^{t}\left(\left\lVert\widetilde{Y_{s}}(\delta(x)\textsc{E}(f))u\right\rVert^{2}+\left\lVert Y_{s}({\langle f(s),\delta(x)\rangle})\textsc{E}(f)u\right\rVert^{2}\right)ds (26)

There’s the following characterization for an integrable map-valued process generated by the structure maps =/2,δ{\mathcal{L}}=-\mathop{}\!\mathbin{\bigtriangleup}/2,\delta from Θ\Theta through the Picard iteration scheme, the convergence of which will yields the solution to the qsde needed.

Lemma 4.2.

[14, lemma 8.1.37] Let 𝒱={(d-valued simple functions}{\mathcal{V}}=\{(d{\mathcal{F}}\text{-valued simple functions}\}, E(𝒱)\textsc{E}({\mathcal{V}}) the exponential vectors, and J(0):E(𝒱)𝒜Γ(k0)J^{(0)}:{\mathcal{F}}\otimes\textsc{E}({\mathcal{V}})\to{\mathcal{A}}\otimes\operatorname{{\Gamma}}(k_{0}) be the identity map, then with J(0)=1J^{(0)}={\textbf{1}},

,J(n+1)(t)=0tJ(n)(s)(aδ+aδ+I)(ds),J(n+1):E(𝒱)𝒜Γ(k0)\displaystyle,J^{(n+1)}(t)=\int_{0}^{t}J^{(n)}(s)\circ(a^{\dagger}_{\delta}+a_{\delta}+I_{\mathcal{L}})(ds),\ \ J^{(n+1)}:{\mathcal{F}}\otimes\textsc{E}({\mathcal{V}})\to{\mathcal{A}}\otimes\operatorname{{\Gamma}}(k_{0}) (27)

each JnJ^{n} is an a map-valued integrable process (by definition linear, but not necessarily completely smooth), Additionally, the following estimates hold ,

Jt(n+1)(xE(f))u2\displaystyle\left\lVert J^{(n+1)}_{t}(x\otimes\textsc{E}(f))u\right\rVert^{2} 2(0tJs(n)(aδ+I)(ds)(xE(f))u2+0tJs(n)(aδ)(ds)(xE(f))u2)\displaystyle\leq 2\left(\left\lVert\int_{0}^{t}J^{(n)}_{s}\circ(a_{\delta}+I_{\mathcal{L}})(ds)(x\otimes\textsc{E}(f))u\right\rVert^{2}+\left\lVert\int_{0}^{t}J^{(n)}_{s}\circ(a^{\dagger}_{\delta})(ds)(x\otimes\textsc{E}(f))u\right\rVert^{2}\right) (28)
Proof.

The continuity requirements for existence of the integral are satisfied since for each fixed E(f)\textsc{E}(f) and aa, the maps δ,\delta,\mathop{}\!\mathbin{\bigtriangleup} are bounded. The inequalities follow from standard theory. The Laplacian flow generator satisfies much weaker assumptions, therefore the resultant process is not completely smooth. ∎

The Picard iterates defined by SN(t)=nNJt(n)(xE(f))S_{N}(t)=\sum_{n\leq N}J_{t}^{(n)}(x\otimes\textsc{E}(f)) can be shown to converge on the exponential vectors following same scheme as [14, Thm 8.1.38] after plugging in the following estimates which need to be obtained differently as Θ\Theta has much less regularity. To motivate the estimates we sketch the convergence arguments.

The first term in r.h.s.for equation 28, using the definition of map-valued integrals (see appendix B42) can be recursively expanded using via estimate 4.1, inequality  25:

et0tJs(n)((x)+δ(x),f(s))E(f))u2ds=et0t0sJs1(n1)(aδ+I)(ds1)(xE(f))u2dse^{t}\int_{0}^{t}\left\lVert J^{(n)}_{s}({\mathcal{L}}(x)+{\langle\delta(x^{*}),f(s)\rangle})\otimes\textsc{E}(f))u\right\rVert^{2}ds=e^{t}\int_{0}^{t}\left\lVert\int_{0}^{s}J^{(n-1)}_{s^{1}}\circ(a_{\delta}+I_{\mathcal{L}})(ds^{1})(x\otimes\textsc{E}(f))u\right\rVert^{2}ds

Since ff is simple, all terms depending on ff in above can be uniformly bound by a constant BfB_{f}, and recursively applying inequality 25 to the r.h.s., till reaching J(0)=1J^{(0)}={\textbf{1}} yields:

J(n+1)(t)(aEf)2\displaystyle\left\lVert J^{(n+1)}(t)(a\otimes\textsc{E}{f})\right\rVert^{2} (2etB)nE(f)20t0s0s1Φf^(s)(Φf^(s1)Φf^(sn1)(x))𝑑s𝑑s1𝑑sn1\displaystyle\leq(2e^{t}B)^{n}\left\lVert\textsc{E}(f)\right\rVert^{2}\int_{0}^{t}\int_{0}^{s}\int_{0}^{s_{1}}\dots\left\lVert\Phi_{\hat{f}(s)}(\Phi_{\hat{f}(s_{1})}\dots\Phi_{\hat{f}(s_{n-1})}(x))\right\rVert dsds_{1}\dots ds_{n-1}

where Φf(s)(x):=(x)+δ(x),f(s)\Phi_{f(s)}(x):={\mathcal{L}}(x)+{\langle\delta(x^{*}),f(s)\rangle}. Since ff is simple, Range[f]\operatorname{{\textsc{Range}}}[f] is finite: for each s[0,t]s\in[0,t], f(s){dζid,i[r],r}Range[f]f(s)\in\{d\zeta_{i}\in d{\mathcal{F}},i\in[r],r\in{\mathds{N}}\}\equiv\operatorname{{\textsc{Range}}}[f]. This means one must control

Φξin(Φξin1(Φξi1(x)))\displaystyle\left\lVert\Phi_{\xi_{i_{n}}}(\Phi_{\xi_{i_{n-1}}}(\dots\Phi_{\xi_{i_{1}}}(x)))\right\rVert (30)

where each ξikRange[f]\xi_{i_{k}}\in\operatorname{{\textsc{Range}}}[f]. Similarly, for the second term in r.h.s.for equation 28, from equation 26 (also, see appendix equation 44), we have

0tJs(n)(aδ)(ds)(xE(f))u2et0t(J(n)~(δ(x)E(f))u2+J(n)(f(s),δ(x))E(f)u2)𝑑s\left\lVert\int_{0}^{t}J^{(n)}_{s}\circ(a^{\dagger}_{\delta})(ds)(x\otimes\textsc{E}(f))u\right\rVert^{2}\leq e^{t}\int_{0}^{t}\left(\left\lVert\widetilde{J^{(n)}}(\delta(x)\otimes\textsc{E}(f))u\right\rVert^{2}+\left\lVert J^{(n)}({\langle f(s),\delta(x)\rangle})\textsc{E}(f)u\right\rVert^{2}\right)ds

The J(n)(f(s),δ(x))E(f)u2\|J^{(n)}({\langle f(s),\delta(x)\rangle})\textsc{E}(f)u\|^{2} term is controlled exactly as equation 4 via a corresponding estimate on nested Φf(s)(x):=f(s),δ(x)\Phi^{\prime}_{f(s)}(x):={\langle f(s),\delta(x)\rangle}. The J(n)~(δ(x)E(f))u2\|\widetilde{J^{(n)}}(\delta(x)\otimes\textsc{E}(f))u\|^{2} term lives on the Fock spaces it can be controlled by same recursive expansion using Φf(s)′′(x):=δ(x)E(f(s))\Phi^{\prime\prime}_{f(s)}(x):=\delta(x)\otimes\textsc{E}(f(s)) since by definition (see appendix equation 43 and remark B.1), integral with respect to aδa^{\dagger}_{\delta} is given by –

(0tJ(n)(s)(aδ)(ds))(xE(f))u\displaystyle\left(\int_{0}^{t}J^{(n)}(s)\circ(a^{\dagger}_{\delta})(ds)\right)(x\otimes\textsc{E}(f))u =0t(Jn(s)~(δ(x)E(fs))u)𝑑s\displaystyle=\int_{0}^{t}\left(\widetilde{J^{n}(s)}(\delta(x)\otimes\textsc{E}(f_{s}))u\right)ds
=0t((0sJ(n1)(s)(aδ)(ds1))(δ(x)E(fs)))u𝑑s\displaystyle=\int_{0}^{t}\left(\left(\int_{0}^{s}J^{(n-1)}(s)\circ(a^{\dagger}_{\delta})(ds_{1})\right)(\delta(x)\otimes\textsc{E}(f_{s}))\right)uds
=0t(0sJ(n1)~((δ1)(δ(x)E(fs)))u)𝑑s1𝑑s\displaystyle=\int_{0}^{t}\left(\int_{0}^{s}\widetilde{J^{(n-1)}}((\delta\otimes{\textbf{1}})(\delta(x)\otimes\textsc{E}(f_{s})))u\right)ds_{1}ds

so one needs to bound both of the following

Φξin(Φξin1(Φξi1(x))),δn(x)\displaystyle\left\lVert\Phi^{\prime}_{\xi_{i_{n}}}(\Phi^{\prime}_{\xi_{i_{n-1}}}(\dots\Phi^{\prime}_{\xi_{i_{1}}}(x)))\right\rVert,\ \left\lVert\delta^{n}(x)\right\rVert (31)
Estimate 4.3.

There exist constants, Cf,x,Mf,xC_{f,x},M_{f,x} such that

i)\displaystyle i)\ Φξin(Φξin1(Φξi1(x)))Cf,xMf,xnii)Φξin(Φξin1(Φξi1(x)))Cf,xMf,xniii)δm(x)Cf,xMf,xn\displaystyle\left\lVert\Phi_{\xi_{i_{n}}}(\Phi_{\xi_{i_{n-1}}}(\dots\Phi_{\xi_{i_{1}}}(x)))\right\rVert\leq C_{f,x}M_{f,x}^{n}\ \ \ ii)\ \left\lVert\Phi^{\prime}_{\xi_{i_{n}}}(\Phi^{\prime}_{\xi_{i_{n-1}}}(\dots\Phi^{\prime}_{\xi_{i_{1}}}(x)))\right\rVert\leq C_{f,x}M_{f,x}^{n}\ \ \ iii)\ \left\lVert\delta^{m}(x)\right\rVert\leq C_{f,x}M_{f,x}^{n}

That is, three terms (equations 30,31) satisfy bound of form CKnCK^{n} for C,KC,K depending on the simple function ff and independent from nn (note the constants for Φ,Φ,Φ′′\Phi,\Phi^{\prime},\Phi^{\prime\prime} may be different, but taking the maximum, yields a single pair that works).

Proof.

i)i) Define an:=Φξin(Φξin1(Φξi1(x)))a_{n}:=\Phi_{\xi_{i_{n}}}(\Phi_{\xi_{i_{n-1}}}(\dots\Phi_{\xi_{i_{1}}}(x))). Since each ξikd\xi_{i_{k}}\in d{\mathcal{F}}, let ξik=dzk\xi_{i_{k}}=dz_{k}. Now Φzk(a)=a+da,dzk=a+a,zk\Phi_{z_{k}}(a^{\prime})=\mathop{}\!\mathbin{\bigtriangleup}^{\prime}a^{\prime}+{\langle da^{\prime},dz_{k}\rangle}=\mathop{}\!\mathbin{\bigtriangleup}^{\prime}a^{\prime}+{\langle\nabla a^{\prime},\nabla z_{k}\rangle} where :=/2\mathop{}\!\mathbin{\bigtriangleup}^{\prime}:=-\mathop{}\!\mathbin{\bigtriangleup}/2. Therefore,

an=Φzn(an1)\displaystyle\left\lVert a_{n}\right\rVert=\Phi_{z_{n}}(a_{n-1}) =an1+an1,zn=ginjn(12injnan1+inan1jnzn)\displaystyle=\mathop{}\!\mathbin{\bigtriangleup}^{\prime}a_{n-1}+{\langle\nabla a_{n-1},\nabla z_{n}\rangle}=g_{i_{n}j_{n}}\left(-\frac{1}{2}\nabla^{i_{n}}\nabla^{j_{n}}a_{n-1}+\nabla^{i_{n}}a_{n-1}\nabla^{j_{n}}z_{n}\right)
=ginjngin1jn1(12injn(12in1jn1an2+in1an2jn1zn1))\displaystyle=g_{i_{n}j_{n}}g_{i_{n-1}j_{n-1}}\left(-\frac{1}{2}\nabla^{i_{n}}\nabla^{j_{n}}\left(-\frac{1}{2}\nabla^{i_{n-1}}\nabla^{j_{n-1}}a_{n-2}+\nabla^{i_{n-1}}a_{n-2}\nabla^{j_{n-1}}z_{n-1}\right)\right)
+ginjngin1jn1(in(12in1jn1an2+in1an2jn1zn1)jnzn)\displaystyle\ \ \ \ \ \ +g_{i_{n}j_{n}}g_{i_{n-1}j_{n-1}}\left(\nabla^{i_{n}}\left(-\frac{1}{2}\nabla^{i_{n-1}}\nabla^{j_{n-1}}a_{n-2}+\nabla^{i_{n-1}}a_{n-2}\nabla^{j_{n-1}}z_{n-1}\right)\nabla^{j_{n}}z_{n}\right)

Therefore, each recursive expansion of Φzk\Phi_{z_{k}} doubles the number of terms. After nn recursive expansion there are 2n2^{n} terms, consisting of nn copies of the metric, ginjngin1jn1g11j1g_{i_{n}j_{n}}g_{i_{n-1}j_{n-1}}\dots g_{1_{1}j_{1}}, contracting against the covariant derivatives acting on x,z1znx,z_{1}\dots z_{n}. Note that ziz_{i}’s appear in increasing order left to right by the recurence structure, and xx is at the left most, making appearance only at the last step. Let Pn+1[I]P_{n+1}[I] be a partition of set I:={ik,jk:k[n]}I:=\{i_{k},j_{k}:k\in[n]\} into n+1n+1 ordered sets Pn+1={AiI:i[n+1]}P_{n+1}=\{A_{i}\subset I:i\in[n+1]\}, with some AiA_{i}’s possibly empty. For APn+1[I]A\in P_{n+1}[I] with A,A=[a1,a2a|A|]\nabla^{A},A=[a_{1},a_{2}\dots a_{|A|}] denote the operator a1a2a|A|\nabla^{a_{1}}\nabla^{a_{2}}\dots\nabla^{a_{|A|}},and by A\nabla_{A} the corresponding operator with indices lowered. Then after nn recursive expansions, each term corresponds to some partition Pn+1[I]={Ai:i[n+1]}P_{n+1}[I]=\{A_{i}:i\in[n+1]\}. Writing the tensor contraction as an explicit innerproduct,

ginjngin1jn1g11j1A1xA2z1A3z2An+1zn=ginjngin1jn1,A1xA2z1A3z2An+1zng_{i_{n}j_{n}}g_{i_{n-1}j_{n-1}}\dots g_{1_{1}j_{1}}\nabla^{A_{1}}x\nabla^{A_{2}}z_{1}\nabla^{A_{3}}z_{2}\dots\nabla^{A_{n+1}}z_{n}={\langle g_{i_{n}j_{n}}g_{i_{n-1}j_{n-1}},\nabla_{A_{1}}x\nabla_{A_{2}}z_{1}\nabla_{A_{3}}z_{2}\dots\nabla_{A_{n+1}}z_{n}\rangle}

where with empty AiA_{i} meaning the term is missing. Applying Cauchy-Schwarz, and dropping the 1/2-1/2 factors as we are only interested in an upper bound on each term,

ginjngin1jn1,A1xA2z1A3z2An+1zn\displaystyle{\langle g_{i_{n}j_{n}}g_{i_{n-1}j_{n-1}},\nabla_{A_{1}}x\nabla_{A_{2}}z_{1}\nabla_{A_{3}}z_{2}\dots\nabla_{A_{n+1}}z_{n}\rangle} (dimM)n/2A1xA2z1An+1zn,A1xA2z1An+1zn\displaystyle\leq(\dim M)^{n/2}\sqrt{{\langle\nabla_{A_{1}}x\nabla_{A_{2}}z_{1}\dots\nabla_{A_{n+1}}z_{n},\nabla_{A_{1}}x\nabla_{A_{2}}z_{1}\nabla_{A_{n+1}}z_{n}\rangle}}
(dimM)n/2A1x,A1xAn+1zn,An+1zn\displaystyle\leq(\dim M)^{n/2}\sqrt{{\langle\nabla_{A_{1}}x,\nabla_{A_{1}}x\rangle}\dots\langle\nabla_{A_{n+1}}z_{n},\nabla_{A_{n+1}}z_{n}\rangle}
(dimM)n/2(|A1|x)(|A2|)z1(|An+1|zn)\displaystyle\leq(\dim M)^{n/2}\ell(\nabla^{|A_{1}|}x)\ell(\nabla^{|A_{2}|})z_{1}\dots\ell(\nabla^{|A_{n+1}|}z_{n})

For each ϕ{x,zn}Tx,f\phi\in\{x,z_{n}\}\equiv T_{x,f}, there are constants Cϕ,MϕC_{\phi},M_{\phi} such that (kϕ)CϕMϕk\ell(\nabla^{k}\phi)\leq C_{\phi}M_{\phi}^{k}. Wlog, assume each Cϕ=1C_{\phi}=1, by absorbing it into MϕM_{\phi}. Now since each dznRange[f],|Range[f]|=rdz_{n}\in\operatorname{{\textsc{Range}}}[f],|\operatorname{{\textsc{Range}}}[f]|=r, the choice M=maxϕTx,fMϕM=\max_{\phi\in T_{x,f}}M_{\phi} is independent of nn, and using that i[n+1]|Ai|=2n\sum_{i\in[n+1]}|A_{i}|=2n, yields

ginjngin1jn1,A1xA2z1A3z2An+1zn((dimM)1/2M2)n{\langle g_{i_{n}j_{n}}g_{i_{n-1}j_{n-1}},\nabla_{A_{1}}x\nabla_{A_{2}}z_{1}\nabla_{A_{3}}z_{2}\dots\nabla_{A_{n+1}}z_{n}\rangle}\leq((\dim M)^{1/2}M^{2})^{n}

And finally accounting 2n2^{n} terms with above inequality holding for each, yields the needed bound

an(2(dimM)1/2M2)n\left\lVert a_{n}\right\rVert\leq(2(\dim M)^{1/2}M^{2})^{n}

ii)ii) This follows identically to part i)i).

iii)iii) Since δ:𝒜𝒜k0,x1dx\delta:{\mathcal{A}}\to{\mathcal{A}}\otimes k_{0},x\to 1\otimes dx, therefore δ2(x)=(δ1)(δ(x))=(δ1)(1dx)=0\delta^{2}(x)=(\delta\otimes{\textbf{1}})(\delta(x))=(\delta\otimes{\textbf{1}})(1\otimes dx)=0. So the bound is simply

sups[0,t]E(fs)dx\sup_{s\in[0,t]}\left\lVert\textsc{E}(f_{s})\right\rVert\left\lVert dx\right\rVert

With these estimates available, the following results are easily obtained exactly as in [14].

Theorem 4.4.

Suppose aa\in{\mathcal{F}}, and f𝒰f\in{\mathcal{U}} where 𝒰{\mathcal{U}} is the set of simple function taking values in (k0)(k_{0})^{\mathcal{F}}_{\infty} such that ft{ξi:i[r]}f_{t}\in\{\xi_{i}:i\in[r]\} for all tt. Define

J(0)=1,J(n+1)=0tJ(n)(s)(aδ+aδ+I)(ds)\displaystyle J^{(0)}={\textbf{1}},J^{(n+1)}=\int_{0}^{t}J^{(n)}(s)\circ(a^{\dagger}_{\delta}+a_{\delta}+I_{\mathcal{L}})(ds) (32)

where δ,\delta,{\mathcal{L}} from the structure matrix Θ\Theta then

  1. 1.

    J(t)=nJt(n)(aE(f))J(t)=\sum_{n}J_{t}^{(n)}(a\otimes\textsc{E}(f)) converges

  2. 2.

    J(t)=1+0tJ(s)(aδ+aδ+I)(ds)J(t)={\textbf{1}}+\int_{0}^{t}J(s)(a^{\dagger}_{\delta}+a_{\delta}+I_{\mathcal{L}})(ds) holds

  3. 3.

    For all a1,a2,f1,f2a_{1},a_{2}\in{\mathcal{F}},f_{1},f_{2} simple and (k0)(k_{0})^{\mathcal{F}}_{\infty}-valued, v1,v2Hv_{1},v_{2}\in{\textsf{H}},

    Jt(a1Ef1)v1,Jt(a2Ef2)v2\displaystyle{\langle J_{t}(a_{1}\otimes\textsc{E}f_{1})v_{1},J_{t}(a_{2}\otimes\textsc{E}f_{2})v_{2}\rangle} =v1Ef1,Jt(a1a2Ef2)v2\displaystyle={\langle v_{1}\otimes\textsc{E}f_{1},J_{t}(a_{1}^{*}a_{2}\otimes\textsc{E}f_{2})v_{2}\rangle}
    J(t)(1Ef1)v1\displaystyle J(t)({\textbf{1}}\otimes\textsc{E}f_{1})v_{1} =v1Ef1\displaystyle=v_{1}\textsc{E}f_{1}
Proof.

From the estimate 4.3, SN(t)=nNJt(n)(aE(f))S_{N}(t)=\sum_{n\leq N}J_{t}^{(n)}(a\otimes\textsc{E}(f)) is convergent since for each Jt(n)(aE(f))J_{t}^{(n)}(a\otimes\textsc{E}(f)),

Jt(n)(aE(f))Cf,x(2etBMf,x)n/n!\left\lVert J_{t}^{(n)}(a\otimes\textsc{E}(f))\right\rVert\leq C_{f,x}(2e^{t}BM_{f,x})^{n}/n!

This yields part 1. Part 2. is by the definition of the stochastic integral, and 3. is standard theory as in [14, Thm 8.1.38]. ∎

It remains to check that the integral extends from E(𝒰)\textsc{E}({\mathcal{U}}) to Γ(k0)\operatorname{{\Gamma}}(k_{0}). For this we need to extend the squareroot trick using regularity of the flow generator.

4.1 The extended squareroot trick

It follows from quantum Ito formulae that for u,vHu,v\in H, h,f𝒱L2(+,k0)kh,f\in{\mathcal{V}}\subset L^{2}(\mathds{R}_{+},k_{0})\equiv k,

Jt(aE(f))u,Jt(bE(h))v\displaystyle{\langle J_{t}(a\otimes\textsc{E}(f))u,J_{t}(b\otimes\textsc{E}(h))v\rangle} =uE(f),Jt(abE(h))v\displaystyle={\langle u\textsc{E}(f),J_{t}(a^{*}b\otimes\textsc{E}(h))v\rangle} (33)
Jt(1E(f))u\displaystyle J_{t}(1\otimes\textsc{E}(f))u =uE(f)\displaystyle=u\textsc{E}(f) (34)

Define

jtn(a)(vEf):=Jt(n)(aEf)vj^{n}_{t}(a)(v\textsc{E}f):=J^{(n)}_{t}(a\otimes\textsc{E}f)v

so jtnj^{n}_{t} is unital with the factorization property (equation 33), and jtn(a)j^{n}_{t}(a) is a linear operator on a dense subspace 𝒦:=HE(𝒱)HΓ(k){\mathcal{K}}:=H\otimes\textsc{E}({\mathcal{V}})\subset H\otimes\operatorname{{\Gamma}}(k). For any v,fv,f, jtnj_{t}^{n} is bounded pointwise on {\mathcal{F}}.

Proposition 4.5.

For all aa\in{\mathcal{F}} there exists KK such that

Θ(a)KaW2,\left\lVert\Theta(a)\right\rVert\leq K\left\lVert a\right\rVert_{W^{2,\infty}}

where W2,\left\lVert\cdot\right\rVert_{W^{2,\infty}} is the Sobolev norm aW2,=a+a+2(a)\left\lVert a\right\rVert_{W^{2,\infty}}=\left\lVert a\right\rVert+\left\lVert\nabla a\right\rVert+\left\lVert\nabla^{2}(a)\right\rVert

Proof.

For a,ψ,fda,\psi\in{\mathcal{F}},f^{\prime}\in d{\mathcal{F}} with f=df,ff^{\prime}=df,f\in{\mathcal{F}} and ww\in\mathbb{C},

((a)/2,da1da0)(ψwψdf)=(((a)/2)ψw+ψdf,daψwda)\displaystyle\begin{pmatrix}-\mathop{}\!\mathbin{\bigtriangleup}(a)/2&{\langle\cdot,da\rangle}\\ 1\otimes da&0\end{pmatrix}\begin{pmatrix}\psi\otimes w\\ \psi\otimes df\end{pmatrix}=\begin{pmatrix}-(\mathop{}\!\mathbin{\bigtriangleup}(a)/2)\psi\otimes w+\psi\otimes{\langle df,da\rangle}\\ \psi\otimes w\otimes da\end{pmatrix} (35)

Therefore, the inequality Θ(a)4dim(M)aW2,\left\lVert\Theta(a)\right\rVert\leq 4\dim(M)\left\lVert a\right\rVert_{W^{2,\infty}} follows, since

Θ(a)(ψ(wdf))2(a)/22ψw2+da2ψdf2+da2ψw2\left\lVert\Theta(a)(\psi\otimes(w\oplus df))\right\rVert^{2}\leq\left\lVert\mathop{}\!\mathbin{\bigtriangleup}(a)/2\right\rVert^{2}\left\lVert\psi\otimes w\right\rVert^{2}+\left\lVert da\right\rVert^{2}\left\lVert\psi\otimes df\right\rVert^{2}+\left\lVert da\right\rVert^{2}\left\lVert\psi\otimes w\right\rVert^{2}

Now Wk,p(M){\mathcal{F}}\subset W^{k,p}(M) since 𝒞(M),M{\mathcal{F}}\subset{\mathcal{C}}^{\infty}(M),M compact for all p,kp,k. Because Θ(a)KaW2,\left\lVert\Theta(a)\right\rVert\leq K\left\lVert a\right\rVert_{W^{2,\infty}}, if Jt(n1)J^{(n-1)}_{t} is bounded for each tt, then Jt(n)J_{t}^{(n)} is continuous on {\mathcal{F}} wit respect to W2,2W^{2,2}-norm topology. We will use this to show that if Jt(n1)J^{(n-1)}_{t} is bounded, then jt(n)j^{(n)}_{t} is positive on 𝒦{\mathcal{K}}. Then using jtnj^{n}_{t} is positive on 𝒦{\mathcal{K}} it will be checked that for every aa\in{\mathcal{F}}, jtn(a)(𝒦)j^{n}_{t}(a)\in{\mathcal{B}}({\mathcal{K}}) and that it extends from (𝒦){\mathcal{B}}({\mathcal{K}}) to (HΓ(k)){\mathcal{B}}(H\otimes\operatorname{{\Gamma}}(k)). The base case is jt(0)=1(𝒦)j_{t}^{(0)}={\textbf{1}}\in{\mathcal{B}}({\mathcal{K}}) which is obviously positive. Then from jtn:(HΓ(k))j_{t}^{n}:{\mathcal{F}}\to{\mathcal{B}}(H\otimes\operatorname{{\Gamma}}(k)), it extends to jtn:𝒞(M)(HΓ(k))j_{t}^{n}:{\mathcal{C}}(M)\to{\mathcal{B}}(H\otimes\operatorname{{\Gamma}}(k)).

Lemma 4.6.

Suppose Jt(n1)(𝒦)J^{(n-1)}_{t}\in{\mathcal{B}}({\mathcal{K}}), then jtnj_{t}^{n} is a positive map on a,a>0a\in{\mathcal{F}},a>0.

Proof.

Suppose aa\in{\mathcal{F}} is positive. We want to show jtn(a)j_{t}^{n}(a) is positive as well. If a\sqrt{a}\in{\mathcal{F}}, then

u,jtn(a)u=jtn(a)u,jtn(a)u0\displaystyle{\langle u,j^{n}_{t}(a)u\rangle}={\langle j^{n}_{t}(\sqrt{a})u,j^{n}_{t}(\sqrt{a})u\rangle}\geq 0 (36)

for every u𝒦u\in{\mathcal{K}}, hence jt(a)j_{t}(a) is positive.

So assume a\sqrt{a}\not\in{\mathcal{F}} where aa is positive, a(m)>0,mMa(m)>0,m\in M. Since LinSpan()\operatorname{{LinSpan}}({\mathcal{F}}) is dense in 𝒞(M){\mathcal{C}}(M), for any ϵ>0\epsilon>0, there exists ff\in{\mathcal{F}} such that afO(ϵ)\left\lVert\sqrt{a}-f\right\rVert\leq O(\epsilon) meaning af2O(ϵ)\left\lVert a-f^{2}\right\rVert\leq O(\epsilon). Additionally ff can be chosen so af2W2,O(ϵ)\left\lVert a-f^{2}\right\rVert_{W^{2,\infty}}\leq O(\epsilon), so f2f^{2} approximates aa in Sobolev W2,W^{2,\infty}-norm as well.

To see why this is possible note that since a>0,aC(M)a>0,\sqrt{a}\in C^{\infty}(M), therefore, aL2(M)\sqrt{a}\in L^{2}(M), additionally for all kk, k(a)L2(M)\nabla^{k}(\sqrt{a})\in L^{2}(M), with

a=iαiϕi and iλi2kαi=k(a),k(a)L2(M)<\sqrt{a}=\sum_{i}\alpha_{i}\phi_{i}\text{ and }\sum_{i}\lambda_{i}^{2k}\alpha_{i}={\langle\nabla^{k}(\sqrt{a}),\nabla^{k}(\sqrt{a})\rangle}_{L^{2}(M)}<\infty

where αi=ϕi,a\alpha_{i}={\langle\phi_{i},\sqrt{a}\rangle} and Ricci-flatness was used. So the sequence (i=1nαiϕi)n(\sum_{i=1}^{n}\alpha_{i}\phi_{i})_{n\in{\mathds{N}}} converging to a\sqrt{a} in L2(M)L^{2}(M) is a bounded in every Wk,2W^{k,2}. For sufficiently large kk, the embedding Wk,2(M)W2,2(M)W^{k,2}(M)\subset W^{2,2}(M) is compact by Rellich–Kondrachov theorem, that is, (i=1nαiϕi)n(\sum_{i=1}^{n}\alpha_{i}\phi_{i})_{n} has a Cauchy, and so a convergent subsequence; wlog let this subsequence be denoted by the same i=1nαiϕi:=an\sum_{i=1}^{n}\alpha_{i}\phi_{i}:=a_{n}. For (ai)(a_{i}) to be convergent in W2,2(M)W^{2,2}(M), aiL2(M)\left\lVert a_{i}\right\rVert_{L^{2}(M)} must vanish, so the only possible limit is a\sqrt{a}. Now suppose the tail i=nαiϕi\sum_{i=n}^{\infty}\alpha_{i}\phi_{i} does not vanish in W2,W^{2,\infty}. This means for some xMx\in M for some k{0,1,2}k\in\{0,1,2\}, iαikϕi,iαikϕi(x)>0{\langle\sum_{i}\alpha_{i}\nabla^{k}\phi_{i},\sum_{i}\alpha_{i}\nabla^{k}\phi_{i}\rangle}(x)>0. But then by the following argument shows that iαikϕi,iαikϕiL2(U)>0{\langle\sum_{i}\alpha_{i}\nabla^{k}\phi_{i},\sum_{i}\alpha_{i}\nabla^{k}\phi_{i}\rangle}_{L^{2}(U)}>0 contradicting the convergence in W2,2(M)W^{2,2}(M). So a\sqrt{a} can be approximated arbitrarily well in W2,(M)W^{2,\infty}(M).

Claim 4.7.

Suppose a=iαiϕi𝒞(M)a=\sum_{i}\alpha_{i}\phi_{i}\in{\mathcal{C}}^{\infty}(M), then kiαiϕi,kiαiϕi(x)>0{\langle\nabla^{k}\sum_{i}\alpha_{i}\phi_{i},\nabla^{k}\sum_{i}\alpha_{i}\phi_{i}\rangle}(x)>0 for some xMx\in M implies kiαiϕi,kiαiϕiL2(M)>0{\langle\nabla^{k}\sum_{i}\alpha_{i}\phi_{i},\nabla^{k}\sum_{i}\alpha_{i}\phi_{i}\rangle}_{L^{2}(M)}>0. In particular, this holds for k=0k=0.

Proof.

As before in proposition 3.7 by smoothness of aa, this holds for all xUx\in U for some open set UU, and with ψ0\psi\geq 0 compactly supported on UU, ψ>0\psi>0 on an open VUV\subset U,

supxUψ(x)ka,kaL2(M)\displaystyle\sup_{x\in U}\psi(x)\cdot{\langle\nabla^{k}a,\nabla^{k}a\rangle}_{L^{2}(M)} =supxUψ(x)Mka,ka𝑑VgUψkiαiϕi,kiαiϕi𝑑Vg>0\displaystyle=\sup_{x\in U}\psi(x)\cdot\int_{M}{\langle\nabla^{k}a,\nabla^{k}a\rangle}dV_{g}\geq\int_{U}{\langle\psi\nabla^{k}\sum_{i}\alpha_{i}\phi_{i},\nabla^{k}\sum_{i}\alpha_{i}\phi_{i}\rangle}dV_{g}>0

Since ψ\psi is compactly supported and bounded, the claim follows. The k=0,k=1k=0,\nabla^{k}={\textbf{1}} specialization is identical. ∎

Now define

𝒲a={a}{f2:f with af2W2,1/n,n}{\mathcal{W}}_{a}=\{a\}\cup\{f^{2}:f\in{\mathcal{F}}\text{ with }\left\lVert a-f^{2}\right\rVert_{W^{2,\infty}}\leq 1/n,n\in{\mathds{N}}\}

then as Θ(a)KaW2,\left\lVert\Theta(a^{\prime})\right\rVert\leq K\left\lVert a^{\prime}\right\rVert_{W^{2,\infty}} and Jt(n1)J^{(n-1)}_{t} is bounded on 𝒦{\mathcal{K}} by hypothesis, the bound in lemma 4.2, implies norm is continuous map with respect to W2,\left\lVert\cdot\right\rVert_{W^{2,\infty}}-topology on 𝒲a{\mathcal{W}}_{a}:

:𝒲a,ajt(a)\left\lVert\cdot\right\rVert:{\mathcal{W}}_{a}\to\mathds{R},a^{\prime}\to\left\lVert j_{t}(a^{\prime})\right\rVert

If jtn(a)j^{n}_{t}(a) is not positive, then there exists u𝒦u\in{\mathcal{K}} such that u,jtn(a)u<0{\langle u,j^{n}_{t}(a)u\rangle}<0. Since norm is continuous, the map au,jtn(a)ua^{\prime}\to{\langle u,j^{n}_{t}(a^{\prime})u\rangle} is also continuous on 𝒲a{\mathcal{W}}_{a}: by Cauchy-Schwartz inequality, u,jtn(a)uujtn(a)uu2KKaW2,{\langle u,j^{n}_{t}(a^{\prime})u\rangle}\leq\left\lVert u\right\rVert\left\lVert j^{n}_{t}(a^{\prime})u\right\rVert\leq\left\lVert u\right\rVert^{2}K^{\prime}K\left\lVert a^{\prime}\right\rVert_{W^{2,\infty}} where KK^{\prime} depends on uu which we fixed and Jtn1\left\lVert J_{t}^{n-1}\right\rVert. This continuity means u,jtn()u<0{\langle u,j^{n}_{t}(\cdot)u\rangle}<0 on some neighborhood containing aa in 𝒲a{\mathcal{W}}_{a}. However, for any neighborhood UU of aa in 𝒲a{\mathcal{W}}_{a} , wU,waw\in U,w\neq a implies w=f2,fw=f^{2},f\in{\mathcal{F}}, so u,jtn(f2)u0{\langle u,j^{n}_{t}(f^{2})u\rangle}\geq 0 by equation 36. Therefore, jtn(a)j^{n}_{t}(a) must be positive. ∎

Lemma 4.8.

If jtnj^{n}_{t} is a positive map on positive a,a>0a,a>0, then jtn(a)a2\left\lVert j^{n}_{t}(a)\right\rVert\leq\left\lVert a\right\rVert^{2}

Proof.

Let xx\in{\mathcal{F}} so (1+ϵ)xx(1+\epsilon)\left\lVert x\right\rVert-x\in{\mathcal{F}} and positive for any ϵ>0\epsilon>0. Define Φϵ(x):=(1+ϵ)x1x𝒞(M)\Phi_{\epsilon}(x):=\sqrt{(1+\epsilon)\left\lVert x\right\rVert{\textbf{1}}-x}\in{\mathcal{C}}(M). Approximate Φϵ(x)\Phi_{\epsilon}(x) from below by zz\in{\mathcal{F}}. Then jtn(Φϵ(x)2z2)>0j_{t}^{n}(\Phi_{\epsilon}(x)^{2}-z^{2})>0 because Φϵ(x)2z2>0\Phi_{\epsilon}(x)^{2}-z^{2}>0 and jtnj_{t}^{n} is positive. This yields θ,(jtn(Φϵ(x)2)jtn(z2))θ0{\langle\theta,(j_{t}^{n}(\Phi_{\epsilon}(x)^{2})-j_{t}^{n}(z^{2}))\theta\rangle}\geq 0 and we have

θ,jtn(Φϵ(x)2)θθ,jtn(z2)θ0{\langle\theta,j_{t}^{n}(\Phi_{\epsilon}(x)^{2})\theta\rangle}\geq{\langle\theta,j_{t}^{n}(z^{2})\theta\rangle}\geq 0

Now the usual squareroot trick takes over: since jtnj_{t}^{n} is unital,

0jtn(z)θ2\displaystyle 0\leq\left\lVert j_{t}^{n}(z)\theta\right\rVert^{2} =θ,jtn(z2)θθ,jt(n)((1+ϵ)x1x)θ\displaystyle={\langle\theta,j_{t}^{n}(z^{2})\theta\rangle}\leq{\langle\theta,j_{t}^{(n)}((1+\epsilon)\left\lVert x\right\rVert{\textbf{1}}-x)\theta\rangle} (37)
θ,jtn(x)θ\displaystyle{\langle\theta,j_{t}^{n}(x)\theta\rangle} θ,jtn((1+ϵ)x1)θ(1+ϵ)xθ,jtn(1)θ=(1+ϵ)xθ2\displaystyle\leq{\langle\theta,j_{t}^{n}((1+\epsilon)\left\lVert x\right\rVert{\textbf{1}})\theta\rangle}\leq(1+\epsilon)\left\lVert x\right\rVert{\langle\theta,j_{t}^{n}({\textbf{1}})\theta\rangle}=(1+\epsilon)\left\lVert x\right\rVert\left\lVert\theta\right\rVert^{2} (38)

Since ϵ\epsilon was arbitrary, θ,jtn(x)θxθ2{\langle\theta,j_{t}^{n}(x)\theta\rangle}\leq\left\lVert x\right\rVert\left\lVert\theta\right\rVert^{2}. Finally,

jtn(x)θ2=jtn(x)θ,jtn(x)θ=θ,jtn(xx)θxxθ2=x2θ2\displaystyle\left\lVert j^{n}_{t}(x)\theta\right\rVert^{2}={\langle j^{n}_{t}(x)\theta,j^{n}_{t}(x)\theta\rangle}={\langle\theta,j^{n}_{t}(x^{*}x)\theta\rangle}\leq\left\lVert x^{*}x\right\rVert\left\lVert\theta\right\rVert^{2}=\left\lVert x\right\rVert^{2}\left\lVert\theta\right\rVert^{2} (39)

. So jtn(x)x2\left\lVert j_{t}^{n}(x)\right\rVert\leq\left\lVert x\right\rVert^{2}, and the bound on jtn\left\lVert j_{t}^{n}\right\rVert is uniform. ∎

Now from density of ,𝒱{\mathcal{F}},{\mathcal{V}} and 𝒦{\mathcal{K}}, each jtnj^{n}_{t} extends from a map jtn:(𝒦)j_{t}^{n}:{\mathcal{F}}\to{\mathcal{B}}({\mathcal{K}}) to jtn:𝒜(HΓ(k))j_{t}^{n}:{\mathcal{A}}\to{\mathcal{B}}(H\otimes\operatorname{{\Gamma}}(k)). Since SN(t)=n[N]JnS_{N}(t)=\sum_{n\in[N]}J^{n} converges, so does S=limNSNS=\lim_{N\to\infty}S_{N}, and therefore limnjtn\lim_{n\to\infty}\sum j_{t}^{n} is the needed flow. Precisely, we have the following result:

Theorem 4.9.

Following notation from theorem 4.4 from define jt(a)(v1Ef1):=Jt(aEf1)v1j_{t}(a)(v_{1}\textsc{E}f_{1}):=J_{t}(a\otimes\textsc{E}f_{1})v_{1}, then

  1. 1.

    jt:(HE(𝒰))j_{t}:{\mathcal{F}}\to{\mathcal{B}}({\textsf{H}}\otimes\textsc{E}({\mathcal{U}})) is a unital *-homomorphism

  2. 2.

    jtj_{t} extends to jt:(HΓ(k0))j_{t}:{\mathcal{F}}\to{\mathcal{B}}({\textsf{H}}\otimes\operatorname{{\Gamma}}(k_{0}))

  3. 3.

    jtj_{t} extends to jt:𝒜(HΓ(k0))j_{t}:{\mathcal{A}}\to{\mathcal{B}}({\textsf{H}}\otimes\operatorname{{\Gamma}}(k_{0}))

Remark 4.10.

A remark on [14] Frechet structures: Proposition 4.5, along with the growth bounds on (kϕ)\ell(\nabla^{k}\phi) suggests that convergence of the stochastic integrals can approached via a variant of complete smoothness regularity condition introduced by [14]. In absence of the group action, the Frechet space structure on k0k_{0} has to be obtained by other methods – in this setting k\nabla^{k} is the natural candidate for defining the Sobolev norms on dk0d{\mathcal{F}}\subset k_{0}. Some work is required to strap that into the Frechet machinery, since [14] require norms to come from derivations on 𝒜=𝒞(M){\mathcal{A}}={\mathcal{C}}(M), but covariant derivative is a derivation on the tensor bundle. Note that the usual Sobolev norms Wk,\left\lVert\cdot\right\rVert_{W^{k,\infty}} can be used to obtain estimates of the type needed: for instance for estimate 4.3, i)i), using that abWp,aWp,bWp,\left\lVert ab\right\rVert_{W^{p,\infty}}\leq\left\lVert a\right\rVert_{W^{p,\infty}}\left\lVert b\right\rVert_{W^{p,\infty}} yields a slightly weaker bound of form Mf,xn2M_{f,x}^{n^{2}} which is still enough to get convergence.

Remark 4.11.

A remark on [1] construction: The polynomial bounded growth condition that is leveraged for the Laplacian is similar to one obtained by [1]. However, this growth cannot be bound on the entire Hilbert space as operator Θ(a)\Theta(a) is only defined on 𝒞(M){\mathcal{C}}^{\infty}(M), and it becomes necessary to work with a dense subspace and utilize some mild continuity to push the necessary estimates through. The continuity is also needed for the extension of the squareroot trick.

 

Appendix A Laplacian on products

Proposition A.1 (Laplacian on products).

For ϕj,ϕi\phi_{j},\phi_{i},

  1. i

    (ϕiϕj)=ϕi(ϕj)+ϕi(ϕj)+2ϕi,ϕj\mathop{}\!\mathbin{\bigtriangleup}(\phi_{i}\phi_{j})=\phi_{i}\mathop{}\!\mathbin{\bigtriangleup}(\phi_{j})+\phi_{i}\mathop{}\!\mathbin{\bigtriangleup}(\phi_{j})+2{\langle\nabla\phi_{i},\nabla\phi_{j}\rangle}

  2. ii

    For any kk,

    kϕi,kϕj=kϕi,kϕj+kϕi,kϕj+2k+1ϕi,k+1ϕj\displaystyle\mathop{}\!\mathbin{\bigtriangleup}{\langle\nabla^{k}\phi_{i},\nabla^{k}\phi_{j}\rangle}={\langle\nabla^{k}\mathop{}\!\mathbin{\bigtriangleup}\phi_{i},\nabla^{k}\phi_{j}\rangle}+{\langle\nabla^{k}\phi_{i},\nabla^{k}\mathop{}\!\mathbin{\bigtriangleup}\phi_{j}\rangle}+2{\langle\nabla^{k+1}\phi_{i},\nabla^{k+1}\phi_{j}\rangle}
Proof.

With =gabab\mathop{}\!\mathbin{\bigtriangleup}=g^{ab}\nabla_{a}\nabla_{b}, MM Ricci flat, using gij=gji,a(ϕiϕj)=ϕja(ϕi)+a(ϕj)ϕig^{ij}=g^{ji},\nabla_{a}(\phi_{i}\phi_{j})=\phi_{j}\nabla_{a}(\phi_{i})+\nabla_{a}(\phi_{j})\phi_{i} yields:

(ϕiϕj)\displaystyle\mathop{}\!\mathbin{\bigtriangleup}(\phi_{i}\phi_{j}) =gabab(ϕiϕj)=gab(ab(ϕiϕj))=gab(a(ϕib(ϕj)+ϕjb(ϕi)))\displaystyle=g^{ab}\nabla_{a}\nabla_{b}(\phi_{i}\phi_{j})=g^{ab}(\nabla_{a}\nabla_{b}(\phi_{i}\phi_{j}))=g^{ab}(\nabla_{a}(\phi_{i}\nabla_{b}(\phi_{j})+\phi_{j}\nabla_{b}(\phi_{i})))
=f(ϕj)+ϕj(ϕi)+2gaba(ϕi)b(ϕj)=f(ϕj)+ϕj(ϕi)+2ϕi,ϕj\displaystyle=f\mathop{}\!\mathbin{\bigtriangleup}(\phi_{j})+\phi_{j}\mathop{}\!\mathbin{\bigtriangleup}(\phi_{i})+2g^{ab}\nabla_{a}(\phi_{i})\nabla_{b}(\phi_{j})=f\mathop{}\!\mathbin{\bigtriangleup}(\phi_{j})+\phi_{j}\mathop{}\!\mathbin{\bigtriangleup}(\phi_{i})+2{\langle\nabla\phi_{i},\nabla\phi_{j}\rangle}

For =gabab\mathop{}\!\mathbin{\bigtriangleup}=-g^{ab}\nabla_{a}\nabla_{b}, we get 2ϕi,ϕj-2{\langle\nabla\phi_{i},\nabla\phi_{j}\rangle} instead for the mixed term.

To see ii)ii), note that by i)i) and for kk-tensors f,hf,h with multi-indices a,ba,b, gab:=ga1b1gakbkg^{ab}:=g^{a_{1}b_{1}}\dots g^{a_{k}b_{k}}, we have

f,h=gabfahb\displaystyle\mathop{}\!\mathbin{\bigtriangleup}{\langle f,h\rangle}=\mathop{}\!\mathbin{\bigtriangleup}g^{ab}f_{a}h_{b} =gabfahb=gabfahb+gab(fa)hb+2gabgcdcfadhb\displaystyle=g^{ab}\mathop{}\!\mathbin{\bigtriangleup}f_{a}h_{b}=g^{ab}f_{a}\mathop{}\!\mathbin{\bigtriangleup}h_{b}+g^{ab}(\mathop{}\!\mathbin{\bigtriangleup}f_{a})h_{b}+2g^{ab}g^{cd}\nabla_{c}f_{a}\nabla_{d}h_{b}
=f,h+f,h+2f,h\displaystyle={\langle\mathop{}\!\mathbin{\bigtriangleup}f,h\rangle}+{\langle f,\mathop{}\!\mathbin{\bigtriangleup}h\rangle}+2{\langle\nabla f,\nabla h\rangle}

and then since[,]=0[\nabla,\mathop{}\!\mathbin{\bigtriangleup}]=0 as MM is Ricci-flat, as needed

kϕi,kϕj=kϕi,kϕj+kϕi,kϕj+k+1ϕi,k+1ϕj\displaystyle\mathop{}\!\mathbin{\bigtriangleup}{\langle\nabla^{k}\phi_{i},\nabla^{k}\phi_{j}\rangle}={\langle\nabla^{k}\mathop{}\!\mathbin{\bigtriangleup}\phi_{i},\nabla^{k}\phi_{j}\rangle}+{\langle\nabla^{k}\phi_{i},\nabla^{k}\mathop{}\!\mathbin{\bigtriangleup}\phi_{j}\rangle}+{\langle\nabla^{k+1}\phi_{i},\nabla^{k+1}\phi_{j}\rangle}

Appendix B Appendix: Map-valued Evans-Hudson quantum sde’s

Suppose 𝒜0{\mathcal{A}}_{0} is a dense *-subalgebra inside the C*-algebra 𝒜(H){\mathcal{A}}\subset{\mathcal{B}}({\textsf{H}}), k0k_{0} is the noise space with k:=L2(+,k0),kt:=L2((0,t),k0),kt:=L2((t,),k0)k:=L^{2}(\mathds{R}_{+},k_{0}),k_{t}:=L^{2}((0,t),k_{0}),k^{t}:=L^{2}((t,\infty),k_{0}). Γ=Γs(k)\operatorname{{\Gamma}}=\operatorname{{\Gamma}}_{s}(k). Γt:=Γ(L2((0,t),k0),Γt:=Γ(L2((t,),k0)\operatorname{{\Gamma}}_{t}:=\operatorname{{\Gamma}}(L^{2}((0,t),k_{0}),\operatorname{{\Gamma}}^{t}:=\operatorname{{\Gamma}}(L^{2}((t,\infty),k_{0}). Similarly, fk,ft,ftf\in k,f^{t},f_{t} are projections onto kt,kt,ft:=f1[0,t],ft:f1[t,)k^{t},k_{t},f_{t}:=f{\textbf{1}}_{[0,t]},f^{t}:f{\textbf{1}}_{[t,\infty)}. k^0:=K0\hat{k}_{0}:=\mathbb{C}\oplus K_{0}.

We will only work with annihilation and creation processes since the conservation process in the Laplacian generated processes is zero. For a map Lin(D0,Hk0)\text{Lin}(D_{0},{\textsf{H}}\otimes k_{0}), D0denseHD_{0}{\subset_{\text{{dense}}}}{\textsf{H}}, if R(u):=abR(u):=a\otimes b, then for Δ(t,)\Delta\subset(t,\infty), using the mapping k0bb1Δktk_{0}\ni b\to b{\textbf{1}}_{\Delta}\in k_{t}, RtΔR^{\Delta}_{t} is defined by

D0ΓtuψRtΔ(uψ):=aψ(b1Δ)HΓtktD_{0}\otimes\operatorname{{\Gamma}}_{t}\ni u\otimes\psi\to R^{\Delta}_{t}(u\otimes\psi):=a\otimes\psi\otimes(b{\textbf{1}}_{\Delta})\in{\textsf{H}}\otimes\operatorname{{\Gamma}}_{t}\otimes k^{t}

The associated creation process creates k0k_{0} component of RR on interval Δ\Delta:

aR(Δ):=a(RtΔ)a_{R}^{\dagger}(\Delta):=a^{\dagger}(R_{t}^{\Delta})

The corresponding annihilation process is defined by using k0k_{0} component of RR to annihilate:

(D0Γt)ΓtutE(ft)aR(δ)(utE(ft))=((ΔR,f(s)𝑑s)ut)E(ft)(D_{0}\otimes\operatorname{{\Gamma}}_{t})\otimes\operatorname{{\Gamma}}^{t}\ni u_{t}\textsc{E}(f^{t})\to a_{R}(\delta)(u_{t}\textsc{E}(f^{t}))=\left(\left(\int_{\Delta}{\langle R,f(s)\rangle}ds\right)u_{t}\right)\textsc{E}(f^{t})

where R,f(s){\langle R,f(s)\rangle} is viewed as an operator() on H.

The Hudson-Parthasarathy quantum stochastic calculus on Hilbert spaces are set in the the Schröedinger picture of quantum dynamics. Suppose (Ht)t0(H_{t})_{t\geq 0} is a family of linear operators on HΓ{\textsf{H}}\otimes\operatorname{{\Gamma}} with {vftnψt}Dom(Ht)\{vf_{t}^{\otimes n}\psi^{t}\}\subset\operatorname{{Dom}}(H_{t}) for vD1denseH,ftktv\in D_{1}{\subset_{\text{{dense}}}}{\textsf{H}},f_{t}\in k_{t}, ftf_{t} simple, right continuous and valued in Vdensek0,ψtΓtV{\subset_{\text{{dense}}}}k_{0},\psi^{t}\in\operatorname{{\Gamma}}^{t}, with Ht=H^t1ΓtH_{t}=\hat{H}_{t}\otimes{\textbf{1}}_{\operatorname{{\Gamma}}^{t}} for some map H^t:{HE(kt)}Dom(H^t)HΓ\hat{H}_{t}:\{{\textsf{H}}\otimes\textsc{E}(k_{t})\}\supset\operatorname{{Dom}}(\hat{H}_{t})\to{\textsf{H}}\otimes\operatorname{{\Gamma}}, such that supstHs(uE(f))rtu\sup_{-\leq s\leq t}\left\lVert H_{s}(u\textsc{E}(f))\right\rVert\leq\left\lVert r_{t}u\right\rVert for all tt, where rtr_{t} depending on t,ft,f only is a closable map in Lin(D1,H)\operatorname{{Lin}}(D_{1},H^{\prime}) for some Hilbert space HH^{\prime} depending only on ff.

If HtH_{t} is simple, that is, Ht=i=0mHti1[ti,ti+1)(t)H_{t}=\sum_{i=0}^{m}H_{t_{i}}{\textbf{1}}_{[t_{i},t_{i+1})}(t), 0=t0<tm<tm+1=0=t_{0}<\dots t_{m}<t_{m+1}=\infty, then for MM as one of the fundamental processes333The integral with respect to the conservation process is not treated here, but the treatment is analogous aR,aR,t1a_{R},a^{\dagger}_{R},t{\textbf{1}},

0tHsM(ds)=0mHtiM([ti,ti+1)[0,t])\int_{0}^{t}H_{s}M(ds)=\sum_{0}^{m}H_{t_{i}}M([t_{i},t_{i+1})\cap[0,t])

For general processes the integral is obtained as an appropriate limit.

The Heisenberg formalism is captured by operator valued processes: for an adapted, regular process, Y(t):𝒜ΓDom(Y(t))𝒜ΓY(t):{\mathcal{A}}\otimes\operatorname{{\Gamma}}\supset\operatorname{{Dom}}(Y(t))\to{\mathcal{A}}\otimes\operatorname{{\Gamma}}, Y(t)~:𝒜k0ΓtDom(Y(t)~)𝒜Γtk0\widetilde{Y(t)}:{\mathcal{A}}\otimes k_{0}\otimes\operatorname{{\Gamma}}_{t}\supset\operatorname{{Dom}}(\widetilde{Y(t)})\to{\mathcal{A}}\otimes\operatorname{{\Gamma}}_{t}\otimes k_{0} defined by Y(t)~=(Y(t)1k0)Swap23\widetilde{Y(t)}=(Y(t)\otimes{\textbf{1}}_{k_{0}}){\text{Swap}_{23}} where Swap23(a1a2a3)=a1a3a2{\text{Swap}_{23}}(a_{1}\otimes a_{2}\otimes a_{3})=a_{1}\otimes a_{3}\otimes a_{2},

Y(s)~:𝒜k0E(ks)Swap23 𝒜E(ks)k0Y(s)1𝒜Γ(ks)k0\widetilde{Y(s)}:{\mathcal{A}}\otimes k_{0}\otimes\textsc{E}(k_{s})\xrightarrow{{\text{Swap}_{23}}\text{ }}{\mathcal{A}}\otimes\textsc{E}(k_{s})\otimes k_{0}\xrightarrow{Y(s)\otimes{\textbf{1}}}{\mathcal{A}}\otimes\operatorname{{\Gamma}}(k_{s})\otimes k_{0}

The map-valued processes for any δ:𝒜0𝒜k0\delta:{\mathcal{A}}_{0}\to{\mathcal{A}}\otimes k_{0} are defined as follows

aδ(Δ)(ixiE(fi))u\displaystyle a_{\delta}(\Delta)(\sum_{i}x_{i}\otimes\textsc{E}(f_{i}))u :=iaδ(xi)(Δ)(uE(fi)),aδ(Δ)(ixiE(fi))u:=iaδ(xi)(Δ)(uE(fi))\displaystyle:=\sum_{i}a_{\delta(x_{i}^{*})}(\Delta)(u\textsc{E}(f_{i})),\ \ a^{\dagger}_{\delta}(\Delta)(\sum_{i}x_{i}\otimes\textsc{E}(f_{i}))u:=\sum_{i}a_{\delta}^{\dagger}(x_{i})(\Delta)(u\textsc{E}(f_{i})) (40)
I(Δ)(ixiE(fi))u\displaystyle I_{\mathcal{L}}(\Delta)(\sum_{i}x_{i}\otimes\textsc{E}(f_{i}))u :=i|Δ|((xi)u)E(fi)\displaystyle:=\sum_{i}|\Delta|({\mathcal{L}}(x_{i})u)\otimes\textsc{E}(f_{i}) (41)

and the map-valued integrals are defined by: with uH,fLloc4,x𝒜u\in{\textsf{H}},f\in L^{4}_{\text{loc}},x\in{\mathcal{A}},

(0tY(s)(aδ+I)(ds))(xE(f))u\displaystyle\left(\int_{0}^{t}Y(s)\circ(a_{\delta}+I_{\mathcal{L}})(ds)\right)(x\otimes\textsc{E}(f))u =0tY(s)(((x)+δ(x),f(s)E(f))uds\displaystyle=\int_{0}^{t}Y(s)\left(({\mathcal{L}}(x)+{\langle\delta(x^{*}),f(s)\rangle}\otimes\textsc{E}(f)\right)u\ ds (42)
(0tY(s)(aδ)(ds))(xE(f))u\displaystyle\left(\int_{0}^{t}Y(s)\circ(a^{\dagger}_{\delta})(ds)\right)(x\otimes\textsc{E}(f))u =(0taY~,x(ds))uE(f)\displaystyle=\left(\int_{0}^{t}a^{\dagger}_{\widetilde{Y},x}(ds)\right)u\textsc{E}(f) (43)

where

aY~,x(s)(uE(fs)):=Y(s)~(δ(x)E(fs))u\displaystyle a^{\dagger}_{\widetilde{Y},x}(s)(u\textsc{E}(f_{s})):=\widetilde{Y(s)}(\delta(x)\otimes\textsc{E}(f_{s}))u (44)
Remark B.1.

A standard technique involves using iterated integrals arising through Picard iteration, requiring iteration of the map δ:𝒜k0\delta:{\mathcal{A}}\to k_{0} of the form δ(δ(x)):=(δ1)(δ(x)):𝒜𝒜k0k0\delta(\delta(x)):=(\delta\otimes{\textbf{1}})(\delta(x)):{\mathcal{A}}\to{\mathcal{A}}\otimes k_{0}\otimes k_{0}, more generally, δm:𝒜𝒜k0m\delta^{m}:{\mathcal{A}}\to{\mathcal{A}}\otimes k_{0}^{\otimes m}. To this, Y~\widetilde{Y} is extended as follows –

Y(s)~:𝒜(mk0m)E(ks)Swap23 𝒜E(ks)(mk0m)Y(s)1𝒜Γ(ks)(mk0m)\widetilde{Y(s)}:{\mathcal{A}}\otimes\left(\oplus_{m}k_{0}^{\otimes m}\right)\otimes\textsc{E}(k_{s})\xrightarrow{{\text{Swap}_{23}}\text{ }}{\mathcal{A}}\otimes\textsc{E}(k_{s})\otimes\left(\oplus_{m}k_{0}^{\otimes m}\right)\xrightarrow{Y(s)\otimes{\textbf{1}}}{\mathcal{A}}\otimes\operatorname{{\Gamma}}(k_{s})\otimes\left(\oplus_{m}k_{0}^{\otimes m}\right)

To make sense of the creation process on k0mk_{0}^{\otimes m}, one just needs an identification k0mktk_{0}^{\otimes m}\to k_{t}, a natural choice is via a1a2ami[0:m1]ai1i|Δ|+Δa_{1}\otimes a_{2}\dots a_{m}\to\sum_{i\in[0:m-1]}a_{i}{\textbf{1}}_{i|\Delta|+\Delta}, where the tensor product leads to the interpretation as a partition refinement.

Now assume (𝒜)𝒜{\mathcal{L}}({\mathcal{A}}^{\infty})\subset{\mathcal{A}}_{\infty}. With 𝒱0=(k0){\mathcal{V}}_{0}=(k_{0})_{\infty}, define

  • 𝒱t={𝒱0-valued simple functions in kt}{\mathcal{V}}_{t}=\{{\mathcal{V}}_{0}\text{-valued simple functions in }k_{t}\}

  • 𝒱={𝒱0-valued simple functions}{\mathcal{V}}=\{{\mathcal{V}}_{0}\text{-valued simple functions}\}

One defines a map-valued integrable process with respect to aδ,aδ,Ia_{\delta},a^{\dagger}_{\delta},I_{\mathcal{L}} as follows:

Definition B.2.

An integrable completely smooth map-valued process is an adapted process (Y(s))s0:𝒜E(𝒱)𝒜Γ(k)(Y(s))_{s\geq 0}:{\mathcal{A}}_{\infty}\otimes\textsc{E}({\mathcal{V}})\to{\mathcal{A}}\otimes\operatorname{{\Gamma}}(k) such that:

  1. 1.

    For each t0,f𝒱t\geq 0,f\in{\mathcal{V}}, Y(t)(aE(f))(𝒜Γ(k))Y(t)(a\otimes\textsc{E}(f))\in({\mathcal{A}}\otimes\operatorname{{\Gamma}}(k))_{\infty} and the map following map is completely smooth:

    𝒜aΩt,f(a):Y(t)(aE(f))(𝒜Γ(k))\displaystyle{\mathcal{A}}_{\infty}\ni a\to\Omega_{t,f}(a):Y(t)(a\otimes\textsc{E}(f))\in({\mathcal{A}}\otimes\operatorname{{\Gamma}}(k))_{\infty} (45)
  2. 2.

    For every mm\in{\mathds{N}} and fixed Xm:=(𝒜k^0m)X\in{\mathcal{E}}^{m}_{\infty}:=({\mathcal{A}}\otimes\hat{k}_{0}^{\otimes m})_{\infty}, f𝒱f\in{\mathcal{V}}, any separable Hilbert space H{\textsf{H}}^{\prime}, the ampilation Ω~t,f:=Ωt,f1H:(𝒜H)(𝒜ΓH)\widetilde{\Omega}_{t,f}:=\Omega_{t,f}\otimes{\textbf{1}}_{{\textsf{H}}^{\prime}}:({\mathcal{A}}\otimes{\textsf{H}}^{\prime})_{\infty}\to({\mathcal{A}}\otimes\operatorname{{\Gamma}}\otimes{\textsf{H}}^{\prime})_{\infty} is continuous.

  3. 3.

    For every fixed a𝒜,f𝒱a\in{\mathcal{A}}_{\infty},f\in{\mathcal{V}}, Y(s)~:=Ω~t,f\widetilde{Y(s)}:=\widetilde{\Omega}_{t,f},

    Sa(s):HE(𝒱s)uE(fs)Y(s)~(δ(aE(fs)))uHΓsk0\displaystyle S_{a}(s):{\textsf{H}}\otimes\textsc{E}({\mathcal{V}}_{s})\ni u\textsc{E}(f_{s})\to\widetilde{Y(s)}(\delta(a\otimes\textsc{E}(f_{s})))u\in{\textsf{H}}\otimes\operatorname{{\Gamma}}_{s}\otimes k_{0} (46)

    is continuous

  4. 4.

    For every a𝒜,ξ𝒱0a\in{\mathcal{A}}_{\infty},\xi\in{\mathcal{V}}_{0} the map

    sY(s)(((a))+δ(a),ξ))E(f)\displaystyle s\to Y(s)(({\mathcal{L}}(a))+{\langle\delta(a^{*}),\xi\rangle}))\otimes\textsc{E}(f) (47)

    is strongly continuous.

References

  • [1] Alexander CR Belton and Stephen J Wills “An algebraic construction of quantum flows with unbounded generators” In Annales de l’IHP Probabilités et statistiques 51.1, 2015, pp. 349–375
  • [2] Nicole Berline, Ezra Getzler and Michele Vergne “Heat kernels and Dirac operators” Springer, 2003
  • [3] Xu Bin “Derivatives of the spectral function and Sobolev norms of eigenfunctions on a closed Riemannian manifold” In Annals of Global Analysis and Geometry 26 Springer, 2004, pp. 231–252
  • [4] Ali H Chamseddine and Alain Connes “Spectral action for Robertson-Walker metrics” In Journal of High Energy Physics 2012.10 Springer, 2012, pp. 1–30
  • [5] Bennett Chow, Peng Lu and Lei Ni “Hamilton’s Ricci flow” American Mathematical Society, Science Press, 2023
  • [6] Nguyen Viet Dang and Michał Wrochna “Complex powers of the wave operator and the spectral action on Lorentzian scattering spaces” In arXiv preprint arXiv:2012.00712, 2020
  • [7] Agostino Devastato, Maxim Kurkov and Fedele Lizzi “Spectral noncommutative geometry standard model and all that” In International Journal of Modern Physics A 34.19 World Scientific, 2019, pp. 1930010
  • [8] José Figueroa-O’Farrill “Spin Geometry” version:18.05.2017 URL: http://empg.maths.ed.ac.uk/Activities/Spin
  • [9] Sita Gakkhar “Quantum diffusion on almost commutative spectral triples and spinor bundles” In arXiv preprint arXiv:2211.03319, 2022 URL: https://arxiv.org/abs/2211.03319v2
  • [10] Emmanuel Hebey “Nonlinear analysis on manifolds: Sobolev spaces and inequalities: Sobolev spaces and inequalities” American Mathematical Soc., 2000
  • [11] H Blaine Lawson and Marie-Louise Michelsohn “Spin Geometry (PMS-38), Volume 38” Princeton university press, 2016
  • [12] John M Lee “Introduction to Riemannian manifolds” Springer, 2018
  • [13] KR Parthasarathy “An introduction to quantum stochastic calculus” Springer Science & Business Media, 1992
  • [14] Kalyan B Sinha and Debashish Goswami “Quantum stochastic processes and noncommutative geometry” Cambridge University Press, 2007
  • [15] J Tolksdorf and T Ackermann “The generalized Lichnerowicz formula and analysis of Dirac operators.” In Journal für die reine und angewandte Mathematik 1996 Walter de Gruyter, Berlin/New York Berlin, New York, 1996 DOI: https://doi.org/10.1515/crll.1996.471.23
  • [16] Walter D Van Suijlekom “Noncommutative geometry and particle physics” Springer, 2015