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

A BBGKY-like Hierarchy for Quantum Field Theories

Michael H. Updike [email protected]    Joshua W. Burby [email protected] Los Alamos National Laboratory, Los Alamos, New Mexico 87545, USA
Abstract

We present a Hamiltonian method of constructing BBGKY-like hierarchies for quantum field theories. With suitable choices, our method creates a hierarchical system of evolution equations for the k-th order reduced density matrices. These equations can be closed at finite order using methods developed for the classical BBGKY hierarchy to give non-perturbative approximations for the full quantum equations of motion. Classical observables can then be numerically computed from these approximate equations, providing an analytically tractable method of modeling high-energy environments where quantum effects play a pronounced role.

I Introduction

Current models of interacting charged particles rely on approximation methods derived from classical electrodynamics. These schemes enjoy great success in relatively weakly interacting environments; however, their predictive power is wanting at higher energies. Large radiative corrections occur when particles begin emitting photons of energy comparable to their mass blackburn . This is not to mention more exotic corrections from effects such as pair creation paircreation . With pulsed laser facilities currently achieving radiation intensities of over 1023W/cm210^{23}W/cm^{2} plasmachallenges , and next-generation burning plasma experiments set to probe comparable energy scales, more sophisticated models must be developed to achieve predictive capabilities. These models must be semi-classical and therefore derive from quantum electrodynamics.

A common starting part for many classical reduced models is the Bogoliubov–Born–Green–
Kirkwood–Yvon (BBGKY) hierarchy. The BBGKY hierarchy is a way of organizing the equations of motion for the NN particle distribution function as a hierarchical system of equations for the kNk\leq N point functions. Through various means, this hierarchy can be reduced to a small number of equations, providing workable approximations to the exact time evolution of a system.

In MMW , the BBGKY hierarchy was shown to have a Hamiltonian structure. In addition to being an important theoretical revelation, this structure offered insight into how BBGKY hierarchies could be constructed for other Hamiltonian theories. Indeed, in Nparticle , and separately in otherpaper , a BBGKY-like hierarchy was obtained for NN boson quantum mechanics. Using a filtration on quantum operators, as opposed to phase space functions, both papers obtained equations for the kkth order reduced density matrices very similar to those of the classical BBGKY hierarchy.

In this paper, we generalize the constructions of MMW , Nparticle , and otherpaper to arbitrary quantum field theories. Using a filtration of Hermitian polynomials in creation/annihilation, we are able to derive a Hamiltonian BBGKY-like hierarchy for the field-theoretic reduced-density matrices. Unlike both the classical and NN-boson hierarchies, our hierarchy has infinitely many equations. Further, compared to the simple coupling of the other hierarchies, our hierarchy generically couples the kkth order variable to the k+1k+1 through k+4k+4th order variables. Despite these differences, methods to truncate the classical BBGKY hierarchy can be applied to our hierarchy without issue, rendering our equations small in number and thus computable.

To the best of our knowledge, the hierarchy we present is novel. We note, however, that a hierarchy of n-particle Wigner functions was constructed for λϕ4\lambda\phi^{4} theory in scalar . While quite different from the approach we employ here, it remains open whether a connection between our hierarchies exists.

With our work, we hope to kickstart the field of using geometric mechanics and quantum field theory to study high-energy-density, many-body systems. Such a viewpoint seems not only natural but also powerful, as witnessed by the relative brevity of our constructions. With further investigation along these lines, we trust that a much greater understanding of high-energy-density environments can be achieved.

II Overview

In III, we generalize the works of MMW , Nparticle , and otherpaper to provide a framework from which a wide class of quantum field theoretic BBGKY-like hierarchies may be constructed. These hierarchies preserve the Hamiltonian structure of the field theory and take a form almost identical to the classical BBGKY hierarchy. In IV, we specialize our discussion to the hierarchy formed from the filtration of ladder operator polynomials, which we dub the canonical hierarchy. We show that the canonical hierarchy has many desirable properties, making it particularly well-suited to approximation schemes. In particular, relevant observables such as the energy and spectral densities depend only on the lowest-level hierarchy variables and are thus easily computed without recourse to further approximations. We demonstrate this in V. Finally, in VI, we offer several avenues of further work including a method of obtaining Hamiltonian closures to our hierarchy.

III The Hamiltonian Theory of QFT Heirachies

For an arbitrary quantum field theory, we let 𝔤\mathfrak{g} denote the space of hermitian operators. This space forms a lie-algebra under the usual commutator bracket i[,]-i[\cdot,\cdot]. A distinguished subset of the dual space 𝔤\mathfrak{g}^{*} is the space of trace-class, positive-semidefinite, hermitian operators 𝔡\mathfrak{d} dualized under the trace map ρ𝔡:Aρ,ATr(Aρ)\rho\in\mathfrak{d}:A\mapsto\left<\rho,A\right>\coloneqq\text{Tr}(A\rho). By normalizing the trace, we may identify every nonzero element of 𝔡\mathfrak{d} with a density matrix. Provided we restrict to unitary equations of motion, 𝔡\mathfrak{d} can therefore be thought of as the phase space of the quantum theory.

We equip 𝔡\mathfrak{d} with a Poisson structure by defining the bracket {,}:C(𝔡)×C(𝔡)C(𝔡)\{\cdot,\cdot\}:C^{\infty}(\mathfrak{d})\times C^{\infty}(\mathfrak{d})\to C^{\infty}(\mathfrak{d})

{F,G}(ρ)=ρ,i[δFδρ,δGδρ],\{F,G\}(\rho)=\left<\rho,-i\left[\frac{\delta F}{\delta\rho},\frac{\delta G}{\delta\rho}\right]\right>, (1)

where δFδρ\frac{\delta F}{\delta\rho} is the unique element of 𝔤\mathfrak{g} satifying

ddϵ|ϵ=0F(ρ+ϵδρ)=δρ,δFδρ.\frac{d}{d\epsilon}|_{\epsilon=0}F(\rho+\epsilon\delta\rho)=\left<\delta\rho,\frac{\delta F}{\delta\rho}\right>. (2)

Given a Hamiltonian H𝔤H\in\mathfrak{g}, the Hamiltonian flow of HH (viewed as a function on 𝔡\mathfrak{d}) provides the usual quantum mechanical equations of motion. Indeed, if ρ(t):J𝔡\rho(t):J\subset\mathbb{R}\to\mathfrak{d} integrates the Hamiltonian flow, then for any FC(𝔡)F\in C^{\infty}(\mathfrak{d}),

ddtF(ρ(t))=δFδρ,ddtρ(t)={F,H}(ρ(t))=iρ(t),[δFδρ,H]=iTr(δFδρ[ρ(t),H]),\frac{d}{dt}F(\rho(t))=\left<\frac{\delta F}{\delta\rho},\frac{d}{dt}\rho(t)\right>=\{F,H\}(\rho(t))=-i\left<\rho(t),\left[\frac{\delta F}{\delta\rho},H\right]\right>=iTr\left(\frac{\delta F}{\delta\rho}\left[\rho(t),H\right]\right), (3)

where the first equality follows from the chain rule and the last from the cyclic property of the trace. It is easy to see that this equality holds for all FC(𝔡)F\in C^{\infty}(\mathfrak{d}) if and only if

ddtρ(t)=i[ρ(t),H],\frac{d}{dt}\rho(t)=i[\rho(t),H], (4)

which is the desired relation.

Let 𝔤1𝔤2𝔤3𝔤\mathfrak{g}_{1}\subset\mathfrak{g}_{2}\subset\mathfrak{g}_{3}\subset\ldots\subset\mathfrak{g} be a sequence of operator subspaces such that i[𝔤n,𝔤m]𝔤n+m1-i[\mathfrak{g}_{n},\mathfrak{g}_{m}]\subset\mathfrak{g}_{n+m-1} and i𝔤i\bigcup_{i}\mathfrak{g}_{i} is dense in 𝔤\mathfrak{g}. We call such a sequence a filtration. For any filtration, we define

𝔊i=1𝔤i.\mathfrak{G}\coloneqq\bigoplus_{i=1}^{\infty}\mathfrak{g}_{i}. (5)

We equip 𝔊\mathfrak{G} with the lie-bracket

[(Ai),(Bi)]𝔊=i([A1,B1],[A2,B1]+[A1,B2],,n+m=k[An,Bm],).[(A_{i}),(B_{i})]_{\mathfrak{G}}=-i\left([A_{1},B_{1}],[A_{2},B_{1}]+[A_{1},B_{2}],\ldots,\sum_{n+m=k}[A_{n},B_{m}],\ldots\right). (6)

As is the case for any direct sum, 𝔊=i=1𝔤i\mathfrak{G}^{*}=\prod_{i=1}^{\infty}\mathfrak{g}^{*}_{i}. Defining the subspaces
𝔡i={ρ|𝔤i:ρ𝔡}𝔤i\mathfrak{d}_{i}=\{\rho|_{\mathfrak{g}_{i}}:\rho\in\mathfrak{d}\}\subset\mathfrak{g}_{i}^{*}, we have the distinguished subspace of 𝔊\mathfrak{G}^{*}

𝔇i=1𝔡i.\mathfrak{D}\coloneqq\prod_{i=1}^{\infty}\mathfrak{d}_{i}. (7)

This space has the canonical Poisson bracket

{F,G}𝔇((ρi))=(ρi),[δFδ(ρi),δGδ(ρi)]𝔊.\{F,G\}_{\mathfrak{D}}((\rho_{i}))=\left<(\rho_{i}),\left[\frac{\delta F}{\delta(\rho_{i})},\frac{\delta G}{\delta(\rho_{i})}\right]_{\mathfrak{G}}\right>. (8)

Writing δFδ(ρi)=(Fi)𝔊\frac{\delta F}{\delta(\rho_{i})}=(F_{i})\in\mathfrak{G}, this is equivalent to the formula

{F,G}((ρi))\displaystyle\left\{F,G\right\}((\rho_{i})) =iTr(ρ1[F1,G1])iTr(ρ2[F2,G1])iTr(ρ2[F1,G2])iTr(ρ3[F3,G1])+.\displaystyle=-iTr\left(\rho_{1}[F_{1},G_{1}]\right)-iTr(\rho_{2}[F_{2},G_{1}])-iTr(\rho_{2}[F_{1},G_{2}])-iTr(\rho_{3}[F_{3},G_{1}])+\ldots. (9)
=iTr(F1[ρ1,G1])+iTr(F2[ρ2,G1])+iTr(F1[ρ2,G2])+iTr(F3[ρ3,G1])+.\displaystyle=iTr\left(F_{1}[\rho_{1},G_{1}]\right)+iTr(F_{2}[\rho_{2},G_{1}])+iTr(F_{1}[\rho_{2},G_{2}])+iTr(F_{3}[\rho_{3},G_{1}])+\ldots.

Let α:𝔊𝔤\alpha:\mathfrak{G}\to\mathfrak{g} denote the natural inclusion α:(Ai)iAi\alpha:(A_{i})\mapsto\sum_{i}A_{i}. One can easily verify that α\alpha is a lie-algebra homomorphism, and consequently that the dual function α:𝔡𝔇,α:ρ(ρ|𝔤i)\alpha^{*}:\mathfrak{d}\to\mathfrak{D},\,\alpha^{*}:\rho\mapsto(\rho|_{\mathfrak{g}_{i}}) is a Poisson map. If (Hi)𝔊(H_{i})\in\mathfrak{G} is any sequence of operators such that α((Hi))=H\alpha((H_{i}))=H, and ρ(t)\rho(t) evolves according to the flow on 𝔡\mathfrak{d}, then αρ(t)\alpha^{*}\rho(t) integrates the flow of (Hi)(H_{i}). This can be verified directly by checking that, for any FC(𝔇)F\in C^{\infty}(\mathfrak{D}),

ddtF(αρ(t))=(Fi),α(ρ(t))={F,(Hi)}(αρ(t)).\frac{d}{dt}F(\alpha^{*}\rho(t))=\left<(F_{i}),\alpha^{*}(\rho(t))\right>=\left\{F,(H_{i})\right\}(\alpha^{*}\rho(t)). (10)

We may view the restricted dual operators ρ|𝔤i\rho|_{\mathfrak{g}_{i}} as equivalence classes [ρ]i[\rho]_{i} of operators under the quotient ρaiρb\rho_{a}\sim_{i}\rho_{b} if and only if Tr(ρaAi)=ρ(ρbAi)Tr(\rho_{a}A_{i})=\rho(\rho_{b}A_{i}) for all Ai𝔤iA_{i}\in\mathfrak{g}_{i}. Letting ρi[ρ]i\rho_{i}\in[\rho]_{i} and πk:A[A]k\pi_{k}:A\mapsto[A]_{k} be the quotient map, we define [[ρ]i,Ai]πi[ρi,Ai][[\rho]_{i},A_{i}]\coloneqq\pi_{i}[\rho_{i},A_{i}] for any Ai𝔤iA_{i}\in\mathfrak{g}_{i}. Some simple algebra reveals that (10) holds if and only if

t[ρ]i=i[[ρ]i,H1]+i[[ρ]i+1,H2]+i[[ρ]i+2,H3]+.\frac{\partial}{\partial t}[\rho]_{i}=i[[\rho]_{i},H_{1}]+i[[\rho]_{i+1},H_{2}]+i[[\rho]_{i+2},H_{3}]+\ldots. (11)

Provided that HBNH\in B_{N} for some small NN, (11) reveals that the equations for [ρ]i[\rho]_{i} form a hierarchical structure. Furthermore, these equations are Hamiltonian. Truncating or otherwise closing the hierarchy, we obtain approximate equations for [ρ]1,,[ρ]N[\rho]_{1},\ldots,[\rho]_{N}. Approximate expectation values are then obtained for every operator OO such that Tr(Oρ)Tr(O\rho) depends only on the equivalence class [ρ]N[\rho]_{N}.

IV The Canonical Hierarchy

Given some set of operators {Xα}\{X_{\alpha}\}, the sequence of subspaces

Bk+1={kth order Hermitian polynomials in Xα,Xα}B_{k+1}=\{\text{kth order Hermitian polynomials in }X_{\alpha},X_{\alpha}^{\dagger}\} (12)

will always form a filtration of 𝔤\mathfrak{g}. In particular, we will see it is desirable to set {Xα}={a𝐩,I}\{X_{\alpha}\}=\{a_{\mathbf{p},I}\}, the annihilation operators of the quantum field theory. Here, 𝐩\mathbf{p} represents the particle momentum index, and II represents any discrete indices (e.g. species, helicity, etc.). This filtration is particularly suited to approximations since, for any renormalizable theory, we may write that H=i=25HiH=\sum_{i=2}^{5}H_{i} with HiBiBi1H_{i}\in B_{i}-B_{i-1}. Defining the reduced density matrices

Γ(I1,,In,I1,,Im)(m,n)[ρ](𝐩1,,𝐩m;𝐩1,,𝐩n)Tr(ρa𝐩1,I1a𝐩n,Ina𝐩1,I1a𝐩m,Im),\Gamma^{(m,n)}_{(I^{\prime}_{1},\ldots,I^{\prime}_{n},I_{1},\ldots,I_{m})}[\rho](\mathbf{p}_{1},...,\mathbf{p}_{m};\mathbf{p}^{\prime}_{1},...,\mathbf{p}^{\prime}_{n})\coloneqq\text{Tr}(\rho a_{\mathbf{p}^{\prime}_{1},I_{1}^{\prime}}^{\dagger}...a_{\mathbf{p}^{\prime}_{n},I_{n}^{\prime}}^{\dagger}a_{\mathbf{p}_{1},I_{1}}...a_{\mathbf{p}_{m},I_{m}}), (13)

this choice of filtration further allows the equivalence classes [ρ]k[\rho]_{k} to be linearly encoded into a finite collection Γk[ρ]=(Γ(m,n)[ρ]:m+n<k)\Gamma_{k}[\rho]=(\Gamma^{(m,n)}_{\mathcal{I}}[\rho]:m+n<k), say by the map ϵk:[ρ]kΓk[ρ]\epsilon_{k}:[\rho]_{k}\mapsto\Gamma_{k}[\rho]. One easily verifies that ϵk\epsilon_{k} is injective and hence invertible on its image. Since applying ϵk\epsilon_{k} commutes with time derivatives, the equations of motion for Γk\Gamma_{k} can be obtained from (11)\eqref{eq:10} as

tΓi=iϵi[ϵi+11(Γi+1),H2]+iϵi[ϵi+21(Γi+2),H3]+iϵi[ϵi+31(Γi+3),H4]+iϵi[ϵi+41(Γi+4),H5].\frac{\partial}{\partial t}\Gamma_{i}=i\epsilon_{i}[\epsilon^{-1}_{i+1}(\Gamma_{i+1}),H_{2}]+i\epsilon_{i}[\epsilon^{-1}_{i+2}(\Gamma_{i+2}),H_{3}]+i\epsilon_{i}[\epsilon^{-1}_{i+3}(\Gamma_{i+3}),H_{4}]+i\epsilon_{i}[\epsilon^{-1}_{i+4}(\Gamma_{i+4}),H_{5}]. (14)

The power of these equations lies in the fact that they are hierarchical and can thus be closed at finite order using methods originally developed for the classical BBGKY hierarchy. The simplest such method is to set Γi0\Gamma_{i}\equiv 0 for i>Ni>N. In this case, the approximate equations of motion read

tΓ1=iϵi[ϵi+11(Γ2),H2]+iϵi[ϵi+21(Γ3),H3]+iϵi[ϵ41(Γi+3),H4]+iϵi[ϵ51(Γi+4),H5],\displaystyle\frac{\partial}{\partial t}\Gamma_{1}=i\epsilon_{i}[\epsilon^{-1}_{i+1}(\Gamma_{2}),H_{2}]+i\epsilon_{i}[\epsilon^{-1}_{i+2}(\Gamma_{3}),H_{3}]+i\epsilon_{i}[\epsilon^{-1}_{4}(\Gamma_{i+3}),H_{4}]+i\epsilon_{i}[\epsilon^{-1}_{5}(\Gamma_{i+4}),H_{5}], (15)
\displaystyle\;\;\;\;\;\;\;\vdots
tΓN3=iϵi[ϵN21(Γ2),HN2]+iϵi[ϵN11(Γ3),HN1]+iϵi[ϵ41(ΓN),H4],\displaystyle\frac{\partial}{\partial t}\Gamma_{N-3}=i\epsilon_{i}[\epsilon^{-1}_{N-2}(\Gamma_{2}),H_{N-2}]+i\epsilon_{i}[\epsilon^{-1}_{N-1}(\Gamma_{3}),H_{N-1}]+i\epsilon_{i}[\epsilon^{-1}_{4}(\Gamma_{N}),H_{4}],
tΓN3=iϵi[ϵN21(Γ2),H2]+iϵi[ϵN11(Γ3),H3]+iϵi[ϵ41(ΓN),H4],\displaystyle\frac{\partial}{\partial t}\Gamma_{N-3}=i\epsilon_{i}[\epsilon^{-1}_{N-2}(\Gamma_{2}),H_{2}]+i\epsilon_{i}[\epsilon^{-1}_{N-1}(\Gamma_{3}),H_{3}]+i\epsilon_{i}[\epsilon^{-1}_{4}(\Gamma_{N}),H_{4}],
tΓN2=iϵi[ϵN11(Γ2),H2]+iϵi[ϵN1(Γ3),H3],\displaystyle\frac{\partial}{\partial t}\Gamma_{N-2}=i\epsilon_{i}[\epsilon^{-1}_{N-1}(\Gamma_{2}),H_{2}]+i\epsilon_{i}[\epsilon^{-1}_{N}(\Gamma_{3}),H_{3}],
tΓN1=iϵi[ϵN1(Γ2),H2],\displaystyle\frac{\partial}{\partial t}\Gamma_{N-1}=i\epsilon_{i}[\epsilon^{-1}_{N}(\Gamma_{2}),H_{2}],
tΓN=0.\displaystyle\frac{\partial}{\partial t}\Gamma_{N}=0.

A more sophisticated method involves cluster expanding the reduced density matrices into their correlations below the NNth order. This scheme similarly reduces (14)\eqref{eq:13} to NN equations. However, this method represents a superior weak-interaction limit since any relativistic quantum field theory has asymptotically vanishing correlations. In the simplest case of a scalar theory and N=1N=1, this amounts to approximating

Γ(m,n)(𝐩1,,𝐩m;𝐩1,,𝐩n)Γ(1,0)¯(𝐩1)Γ(1,0)¯(𝐩n)Γ(1,0)(𝐩1)Γ(1,0)(𝐩m).\Gamma^{(m,n)}(\mathbf{p}_{1},...,\mathbf{p}_{m};\mathbf{p}^{\prime}_{1},...,\mathbf{p}^{\prime}_{n})\approx\overline{\Gamma^{(1,0)}}(\mathbf{p}_{1}^{\prime})\ldots\overline{\Gamma^{(1,0)}}(\mathbf{p}_{n}^{\prime})\ldots\Gamma^{(1,0)}(\mathbf{p}_{1})\ldots\Gamma^{(1,0)}(\mathbf{p}_{m}). (16)

Substituting this expression into (14)\eqref{eq:13} for i=1i=1, we obtain an equation of the form tΓ1=F(Γ1)\partial_{t}\Gamma_{1}=F(\Gamma_{1}) which can be solved and substituted into (16)\eqref{above}.

A major fault in both of these methods is that they do not produce Hamiltonian equations of motion, and hence break the underlying structure of the quantum theory. Ideally, approximations to (14) should close the hierarchy in a Hamiltonian manner. However, a general method for doing so remains open. In VI, we offer several possible routes toward this goal.

We note that once an approximation method has been chosen, the equations of motion are most effectively computed using the relation

iϵi[ϵj1(Γ(m,n)),A]=iTr(ρ[a𝐩1,I1a𝐩n,Ina𝐩1,I1a𝐩m,Im,A]).i\epsilon_{i}[\epsilon^{-1}_{j}(\Gamma_{\mathcal{I}}^{(m,n)}),A]=-i\text{Tr}(\rho\,[a_{\mathbf{p}^{\prime}_{1},I_{1}^{\prime}}^{\dagger}...a_{\mathbf{p}^{\prime}_{n},I_{n}^{\prime}}^{\dagger}a_{\mathbf{p}_{1},I_{1}}...a_{\mathbf{p}_{m},I_{m}},A]). (17)

For example, consider a free scalar theory with the Hamiltonian

H=d3𝐪(2π)3(2E𝐪)E𝐪a𝐪a𝐪B3,H=\int\frac{d^{3}\mathbf{q}}{(2\pi)^{3}(2E_{\mathbf{q}})}E_{\mathbf{q}}a^{\dagger}_{\mathbf{q}}a_{\mathbf{q}}\in B_{3}, (18)

where E𝐩=𝐩2+m2E_{\mathbf{p}}=\sqrt{\mathbf{p}^{2}+m^{2}} is the particle energy. We work in natural units and assume the commutation relations

[a𝐩,,a𝐩,]=(2π)3(2E𝐩)δI,Iδ(3)(𝐩𝐩).[a_{\mathbf{p},\mathcal{I}},a^{\dagger}_{\mathbf{p^{\prime}},\mathcal{I}^{\prime}}]=(2\pi)^{3}(2E_{\mathbf{p}})\,\delta_{I,I^{\prime}}\,\delta^{(3)}(\mathbf{p}-\mathbf{p}^{\prime}). (19)

A simple calculation reveals that

tΓ(n,m)(t;𝐩1,,𝐩n;𝐩¯1,𝐩¯m)=(ij=1n𝐩jt+ij=1m𝐩¯jt)Γ(n,m),\frac{\partial}{\partial t}\Gamma^{(n,m)}(t;\mathbf{p}_{1},\ldots,\mathbf{p}_{n};\overline{\mathbf{p}}_{1}\ldots,\overline{\mathbf{p}}_{m})=\left(-i\sum_{j=1}^{n}\mathbf{p}_{j}t+i\sum_{j=1}^{m}\overline{\mathbf{p}}_{j}t\right)\Gamma^{(n,m)}, (20)

with similarly trivial behavior for other free theories.

For interacting theories, (14)\eqref{eq:13} is no longer exactly solvable. If the scalar Hamiltonian was instead

H=12d3𝐪(2π)3(2E𝐪)d3𝐬(2π)2(2E𝐬)h(𝐪,𝐬)a𝐪a𝐬a𝐪a𝐬H=\frac{1}{2}\int\frac{d^{3}\mathbf{q}}{(2\pi)^{3}(2E_{\mathbf{q}})}\frac{d^{3}\mathbf{s}}{(2\pi)^{2}(2E_{\mathbf{s}})}h(\mathbf{q},\mathbf{s})a_{\mathbf{q}}^{\dagger}a_{\mathbf{s}}^{\dagger}a_{\mathbf{q}}a_{\mathbf{s}} (21)

with h(𝐪,𝐬)h(\mathbf{q},\mathbf{s}) a symmetric, real-valued function, then a similar computation reveals that

tΓ(1,0)(t;𝐩)\displaystyle\frac{\partial}{\partial t}\Gamma^{(1,0)}(t;\mathbf{p}) =id3𝐪(2π)3(2E𝐪)Γ(2,1)(t;𝐩,𝐪;𝐪)h(𝐩,𝐪),\displaystyle=-i\int\frac{d^{3}\mathbf{q}}{(2\pi)^{3}(2E_{\mathbf{q}})}\;\;\;\Gamma^{(2,1)}(t;\mathbf{p},\mathbf{q};\mathbf{q})h(\mathbf{p},\mathbf{q}), (22)
tΓ(1,1)(t;𝐩;𝐩)\displaystyle\frac{\partial}{\partial t}\Gamma^{(1,1)}(t;\mathbf{p};\mathbf{p}^{\prime}) =id3𝐪(2π)3(2E𝐪)Γ(2,2)(t;𝐩,𝐪;𝐩,𝐪)(h(𝐩,𝐪)h(𝐩,𝐪)),\displaystyle=-i\int\frac{d^{3}\mathbf{q}}{(2\pi)^{3}(2E_{\mathbf{q}})}\;\;\Gamma^{(2,2)}(t;\mathbf{p},\mathbf{q};\mathbf{p}^{\prime},\mathbf{q})\left(h(\mathbf{p},\mathbf{q})-h(\mathbf{p}^{\prime},\mathbf{q})\right),
\displaystyle\vdots

which is no longer trivial, highlighting the general need to seek approximations to (14)\eqref{eq:13}.

V Computing Observables

Perhaps the simplest observable is the expected density of particles with discrete indices II and momentum 𝐩\mathbf{p},

DI(𝐩)1(2π)32E𝐩a𝐩,Ia𝐩,I=1(2π)32E𝐩Γ(I,I)(1,1)(𝐩;𝐩).D_{I}(\mathbf{p})\coloneqq\left<\frac{1}{(2\pi)^{3}2E_{\mathbf{p}}}a_{\mathbf{p},I}^{\dagger}a_{\mathbf{p},I}\right>=\frac{1}{(2\pi)^{3}2E_{\mathbf{p}}}\Gamma_{(I,I)}^{(1,1)}(\mathbf{p};\mathbf{p}). (23)

Another observable of interest is the total energy stored in a region of space. For any relativistic theory, we may write the Hamiltonian as

H=d3𝐱(𝐱).H=\int d^{3}\mathbf{x}\;\mathcal{H}(\mathbf{x}). (24)

where (𝐱)\mathcal{H}(\mathbf{x}) is the (0,0)(0,0) component of the physical stress-energy tensor. The energy distribution of a system is then

(𝐱)(𝐱)=F(Γ5).\mathcal{E}(\mathbf{x})\coloneqq\left<\mathcal{H}(\mathbf{x})\right>=F(\Gamma_{5}).

where FF is some function depending on the Hamiltonian. For interacting theories, (𝐱)\left<\mathcal{H}(\mathbf{x})\right> is in general a very complicated function. In many cases, however, the energy stored in particle interactions is much smaller than the energy of the particles themselves. For such cases, we may approximate

(𝐱)free(𝐱)=Ffree(Γ3),\left<\mathcal{H}(\mathbf{x})\right>\approx\left<\mathcal{H}_{\text{free}}(\mathbf{x})\right>=F_{\text{free}}(\Gamma_{3}), (25)

with free\mathcal{H}_{\text{free}} being the Hamiltonian density of the free theory. For example, for an arbitrary scalar theory, we may use (18)\eqref{eq:17} to approximately compute that

(𝐱)\displaystyle\mathcal{E}(\mathbf{x}) d3𝐤(2π)3(2E𝐤)d3𝐩(2π)3(2E𝐩)\displaystyle\approx\int\frac{d^{3}\mathbf{k}}{(2\pi)^{3}(2E_{\mathbf{k}})}\;\frac{d^{3}\mathbf{p}}{(2\pi)^{3}(2E_{\mathbf{p}})}\; (26)
(m2E𝐩E𝐤+𝐩𝐤)ei𝐱(𝐤+𝐩)Γ(2,0)(t;𝐤,𝐩)+ei𝐱(𝐤𝐩)(m2+E𝐩E𝐤+𝐩𝐤)Γ(1,1)(t;𝐩;𝐤).\displaystyle(m^{2}-E_{\mathbf{p}}E_{\mathbf{k}}+\mathbf{p}\cdot\mathbf{k})e^{i\mathbf{x}\cdot(\mathbf{k}+\mathbf{p})}\Gamma^{(2,0)}(t;\mathbf{k},\mathbf{p})+e^{-i\mathbf{x}\cdot(\mathbf{k}-\mathbf{p})}(m^{2}+E_{\mathbf{p}}E_{\mathbf{k}}+\mathbf{p}\cdot\mathbf{k})\Gamma^{(1,1)}(t;\mathbf{p};\mathbf{k}).

For a massive theory, we may further consider the limit whereby the particle mass is much larger than the particle momenta, and hence that the energy of the fields resides entirely in the mass energy of the particles. Letting A𝐩0\left<A\right>_{\mathbf{p}\to 0} denote the expectation value of an operator AA in such a limit, the number density of particles can thus be computed as

N(𝐱)=free(𝐱)𝐩0m.N(\mathbf{x})=\frac{\left<\mathcal{H}_{\text{free}}(\mathbf{x})\right>_{\mathbf{p}\to 0}}{m}. (27)

In the case of a scalar theory, for example, the expected particle density is

N(𝐱)=d3𝐤(2π)32E𝐤d3𝐩(2π)32E𝐩ei𝐱(𝐤𝐩)Γ(1,1)(t;𝐩;𝐤),\displaystyle N(\mathbf{x})=\int\frac{d^{3}\mathbf{k}}{(2\pi)^{3}\sqrt{2E_{\mathbf{k}}}}\frac{d^{3}\mathbf{p}}{(2\pi)^{3}\sqrt{2E_{\mathbf{p}}}}e^{-i\mathbf{x}\cdot(\mathbf{k}-\mathbf{p})}\Gamma^{(1,1)}(t;\mathbf{p};\mathbf{k}), (28)

which is verified as the expectation value of the operator

𝒩(𝐱)=d3𝐤(2π)32E𝐤d3𝐩(2π)32E𝐩ei𝐱(𝐤𝐩)a𝐤a𝐩.\mathcal{N}(\mathbf{x})=\int\frac{d^{3}\mathbf{k}}{(2\pi)^{3}\sqrt{2E_{\mathbf{k}}}}\frac{d^{3}\mathbf{p}}{(2\pi)^{3}\sqrt{2E_{\mathbf{p}}}}e^{-i\mathbf{x}\cdot(\mathbf{k}-\mathbf{p})}a_{\mathbf{k}}^{\dagger}a_{\mathbf{p}}. (29)

Indeed, the interpretation of 𝒩(𝐱)\mathcal{N}(\mathbf{x}) as a number-density is confirmed upon checking that d3𝐱𝒩(𝐱)\int d^{3}\mathbf{x}\;\mathcal{N}(\mathbf{x}) is the usual number operator. Similar operators for other theories are obtained from (29) by adding the appropriate discrete indices and external leg factors.

VI Discussion

While an important first step, several more steps must be taken before our hierarchy can be used in the semi-classical modeling of high-energy environments. For one, before any computational work is undertaken, (14)\eqref{eq:13} must be recast as a differential equation and the desired initial conditions rewritten as reduced density matrices. This is straightforward in the simplest cases but requires considerable effort for realistic theories such as quantum electrodynamics. A more subtle issue is how nontrivial boundary conditions may be worked into our theory. It is clear this involves working in position space, but we leave the details to future work.

A more theoretical line of future work lies in finding Hamiltonian closures to our hierarchy equations. While it is desirable that approximations to the quantum evolution equations preserve the Hamiltonian structure of theory, it is not clear how this can be done. One promising line of investigation is to look for Poisson maps from some smaller Poisson manifold PP to 𝔇\mathfrak{D}. Solutions to the Hamiltonian flow on PP would then push forward to solutions on 𝔇\mathfrak{D}, while presumably being easier to solve. Given an operator ideal \mathcal{I}, a natural choice for PP would be the subspace of 𝔇\mathfrak{D} consisting of sequences (ρ1,ρ2,)(\rho_{1},\rho_{2},\ldots) such that ρi(𝔤i)=0\rho_{i}(\mathfrak{g}_{i}\cap\mathcal{I})=0, with the Poisson bracket inherited from the quotient algebra 𝔤/\mathfrak{g}/\mathcal{I}. However, such ideals remain elusive if they exist.

In a similar vein, one could look for Poisson structures on i=1N𝔡i\prod_{i=1}^{N}\mathfrak{d}_{i} close to the Poisson structure on i=1𝔡i\prod_{i=1}^{\infty}\mathfrak{d}_{i}. Although no longer an exact solution, solving for the flow on the smaller space would still serve as a Hamiltonian approximation to the full equations of motion.

VII Acknowledgment

Research presented in this article was supported by the Los Alamos National Laboratory LDRD program under project number 20230497ECR.

LA-UR Number 23-31725

References

  • [1] T. Blackburn. Radiation reaction in electron–beam interactions with high-intensity lasers. Reviews of Modern Plasma Physics, 4, 2020.
  • [2] N. Bogoliubov. Lectures on Quantum Statistics: Vol. 1. Macdonald technical and scientific, 1967.
  • [3] G. Borgioli and V. Gerasimenko. On the initial-value problem to the quantum dual bbgky hierarchy. Il Nuovo Cimento C, 33, 06 2010.
  • [4] F. Buchholz, I. Theophilou, S. E. Nielsen, M. Ruggenthaler, and A. Rubio. Reduced density-matrix approach to strong matter-photon interaction. ACS photonics, 6(11):2694–2711, 2019.
  • [5] J. W. Burby and P. J. Morrison. Mean field theory for intense light-matter interactions in high energy density plasmas, 2022.
  • [6] V. M. Galitskii and A. B. Migdal. Application of quantum field theory methods to the many body problem. Sov. Phys. JETP, 7(96):18, 1958.
  • [7] V. Guillemin and S. Sternberg. The moment map and collective motion. Annals of Physics, 127(1):220–253, 1980.
  • [8] R. Janev, S. Zhang, and J. Wang. Review of quantum collision dynamics in debye plasmas. Matter and Radiation at Extremes, 1(5):237–248, 2016.
  • [9] Y.-D. Jung. Quantum-mechanical effects on electron–electron scattering in dense high-temperature plasmas. Physics of Plasmas, 8(8):3842–3844, 2001.
  • [10] V. Kaplunovsky. Qft supplementary notes. https://web2.ph.utexas.edu/~vadim/Classes/2022f/notes.html.
  • [11] S. Krönke and P. Schmelcher. Born-bogoliubov-green-kirkwood-yvon hierarchy for ultracold bosonic systems. Phys. Rev. A, 98:013629, Jul 2018.
  • [12] C.-K. Li and R. D. Petrasso. Charged-particle stopping powers in inertial confinement fusion plasmas. Physical review letters, 70(20):3059, 1993.
  • [13] K. A. Lyakhov. Bbgky hierarchy in scalar qft, 2001.
  • [14] V. A. Malyshev. Cluster expansions in lattice models of statistical physics and the quantum theory of fields. Russian Mathematical Surveys, 35(2):1, apr 1980.
  • [15] G. Manfredi. How to model quantum plasmas. Fields Inst. Commun, 46:263–287, 2005.
  • [16] J. Marsden, T. Ratiu, and A. Weinstein. Semidirect products and reduction in mechanics. Transactions of the American Mathematical Society, 281, 01 1984.
  • [17] J. E. Marsden, P. J. Morrison, and A. Weinstein. The Hamiltonian structure of the BBGKY hierarchy equations. In Fluids and plasmas: geometry and dynamics (Boulder, Colo., 1983), volume 28 of Contemp. Math., pages 115–124. Amer. Math. Soc., Providence, RI, 1984.
  • [18] D. Mendelson, A. R. Nahmod, N. Pavlović, M. Rosenzweig, and G. Staffilani. A rigorous derivation of the hamiltonian structure for the nonlinear schrödinger equation, 2019.
  • [19] P. Morrison and F. Pegoraro. On the hamiltonian structure of the bbgky hierarchy. In APS Division of Plasma Physics Meeting Abstracts, volume 2021, pages UP11–040, 2021.
  • [20] P. J. Morrison. Structure and structure-preserving algorithms for plasma physics. Physics of Plasmas, 24(5), 2017.
  • [21] Y. N. Orlov and I. Pavlotsky. Quantum bbgky-hierarchies and wigner’s equation in postgalilean approximation. Physica A: Statistical Mechanics and its Applications, 158(2):607–618, 1989.
  • [22] M. Perin, C. Chandre, P. Morrison, and E. Tassi. Higher-order hamiltonian fluid reduction of vlasov equation. Annals of Physics, 348:50–63, 2014.
  • [23] R. Rajaraman. Some non-perturbative semi-classical methods in quantum field theory (a pedagogical review). Physics Reports, 21(5):227–313, 1975.
  • [24] R. Requist. Hamiltonian formulation of nonequilibrium quantum dynamics: Geometric structure of the bogoliubov-born-green-kirkwood-yvon hierarchy. Physical Review A, 86(2):022117, 2012.
  • [25] R. Requist. Hamiltonian formulation of nonequilibrium quantum dynamics: Geometric structure of the bogoliubov-born-green-kirkwood-yvon hierarchy. Phys. Rev. A, 86:022117, Aug 2012.
  • [26] S. Rose. Electron–positron pair creation in burning thermonuclear plasmas. High Energy Density Physics, 9(3):480–483, 2013.
  • [27] S. J. Rose, P. W. Hatfield, and R. H. H. Scott. Modelling burning thermonuclear plasma. Philosophical Transactions of the Royal Society A: Mathematical, Physical and Engineering Sciences, 378(2184):20200014, 2020.
  • [28] G. Schmidt. Physics of high temperature plasmas. Elsevier, 2012.
  • [29] C. Scovel and A. Weinstein. Finite dimensional lie-poisson approximations to vlasov-poisson equations. Communications on Pure and Applied Mathematics, 47(5):683–709, 1994.
  • [30] S. Weinberg. The Quantum Theory of Fields, volume 1. Cambridge University Press, 1995.