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

On nuclear short-range correlations and the zero-energy eigenstates of the Schrödinger equation

Saar Beck The Racah Institute of Physics, The Hebrew University, Jerusalem 9190401, Israel    Ronen Weiss Theoretical Division, Los Alamos National Laboratory, Los Alamos, New Mexico 87545, USA    Nir Barnea The Racah Institute of Physics, The Hebrew University, Jerusalem 9190401, Israel
Abstract

We present a systematic analysis of the nuclear 2 and 3-body short range correlations, and their relations to the zero-energy eigenstates of the Schrödinger equation. To this end we analyze the doublet and triplet Coupled-Cluster amplitudes in the high momentum limit, and show that they obey universal equations independent of the number of nucleons and their state. Furthermore, we find that these Coupled-Cluster amplitudes coincide with the zero-energy Bloch-Horowitz operator. These results illuminate the relations between the nuclear many-body theory and the generalized contact formalism, introduced to describe the nuclear 2-body short range correlations, and it might also be helpful for general Coupled-Cluster computations as the asymptotic part of the amplitudes is given and shown to be universal.

preprint: LA-UR-22-31633

I Introduction

Nuclear short-range correlations (SRC) have been studied extensively over the last few decades (see Refs. [1, 2] for recent reviews). Large momentum-transfer quasi-elastic electron and proton scattering reactions are the main experimental tools facilitating these studies [3, 4]. In such reactions, interpreted in a high resolution picture, back-to-back SRC nucleon pairs were clearly identified [5, 6, 7, 8, 9, 10], with a significant dominance of neutron-proton pairs [11, 12, 9, 13, 14]. Inclusive reactions where used to study the abundance of such SRC pairs [15, 16, 17, 18, 19]. Currently, ab-initio approaches are unable to directly calculate the cross sections of these reactions, in all but the lightest nuclei. Nevertheless, qualitatively similar conclusions were obtained in structure studies, that focused mainly on the high momentum tail of the nuclear momentum distribution [20, 21, 22, 23, 24, 25, 22, 26, 27, 28]. The study of nuclear three-body SRCs, i.e. three nucleons at close proximity, is still very preliminary at this stage [29, 30] and their impact on nuclear quantities is still mostly unknown.

Following Tan’s work on ultra-cold atoms [31, 32, 33, 34], the Generalized Contact Formalism (GCF) was introduced and utilized to analyze SRC effects in nuclei [35, 36, 37, 38]. It is based on the scale separation ansatz, assuming a factorization of the nuclear wave-function when two nucleons are close to each other. The GCF provides a framework to study both nuclear structure and nuclear reactions, and was successfully tested against ab-initio studies, providing a good description of both two-body densities at short distance and high-momentum tails of different momentum distributions [36, 39, 40]. In addition, the GCF is found to be in good agreement with exclusive electron scattering experiments and other reactions sensitive to SRC pairs [37, 41, 42, 13, 43, 12, 13, 44, 14, 45]. As such, the GCF allows for a quantitative comparison between ab-initio calculations and experimental results, with direct connection to the underlying nuclear interaction. The GCF results lead to a comprehensive and consistent picture of nuclear SRCs, where the only tension is with respect to the analysis of inclusive reactions [46]. Recently, shell-model calculations were combined with the GCF to calculate nuclear matrix elements for neutrinoless double beta decay [47], taking into account both short-range and long-range contributions consistently.

As pointed out, the GCF is based on the asymptotic factorization ansatz for the many-body nuclear wave-function Ψ\Psi, when nucleon ii is close to nucleon jj [36]

Ψrij0αφijα(𝒓ij)Aijα(𝑹ij,{𝒓k}ki,j).\Psi\xrightarrow[r_{ij}\rightarrow 0]{}\sum_{\alpha}\varphi_{ij}^{\alpha}\left(\boldsymbol{r}_{ij}\right)A_{ij}^{\alpha}\left(\boldsymbol{R}_{ij},\{\boldsymbol{r}_{k}\}_{k\not=i,j}\right). (1)

In this picture, particles ii and jj are strongly interacting, and, therefore, described by a two-body function φijα\varphi_{ij}^{\alpha}, decoupled from the rest of the system, which is described by the function AijαA_{ij}^{\alpha}. In the GCF, φijα\varphi_{ij}^{\alpha} is assumed to be universal, i.e. independent of the nucleus or its many-body state, and is defined to be the zero-energy solution of the two-body Schrödinger equation with quantum numbers α\alpha, obtained with the same nucleon-nucleon interaction model used for the many-body wave-function. A similar factorization should hold in momentum space, for pairs with high relative momentum 𝒌ij\boldsymbol{k}_{ij}

Ψ~(𝒌1,𝒌2,,𝒌A)kijαφ~ijα(𝒌ij)A~ijα(𝑲ij,{𝒌n}ni,j),\displaystyle\tilde{\Psi}(\boldsymbol{k}_{1},\boldsymbol{k}_{2},...,\boldsymbol{k}_{A})\xrightarrow[k_{ij}\rightarrow\infty]{}\sum_{\alpha}\tilde{\varphi}_{ij}^{\alpha}\left(\boldsymbol{k}_{ij}\right)\tilde{A}_{ij}^{\alpha}\left(\boldsymbol{K}_{ij},\{\boldsymbol{k}_{n}\}_{n\not=i,j}\right), (2)

where φ~ijα\tilde{\varphi}_{ij}^{\alpha} and A~ijα\tilde{A}_{ij}^{\alpha} are respectively the Fourier transforms of φijα\varphi_{ij}^{\alpha} and AijαA_{ij}^{\alpha}. Based on these asymptotic factorizations, nuclear contact matrices are defined as

Cijαβ=NijAijα|Aijβ.C_{ij}^{\alpha\beta}=N_{ij}\langle A_{ij}^{\alpha}|A_{ij}^{\beta}\rangle. (3)

Here, ijij stands for one of the pairs: proton-proton, neutron-neutron or neutron-proton, and NijN_{ij} is the total number of ijij pairs in the nucleus. The diagonal contact elements CijααC_{ij}^{\alpha\alpha} are proportional to the number of SRC pairs with qunatum number α\alpha in a given nuclear state.

The asymptotic factorization, including the definition of the universal two-body functions, is the underlying assumption for the GCF predictions, and was verified numerically using ab-initio calculations [36, 39, 40]. It is also supported by the work of Refs. [48, 49, 50], based on renormalization group arguments. In view of its success, the two-body GCF is expected to be the leading order term of a short-range (or a high-momentum) expansion of the nuclear wave-function. However, next order corrections are currently not well understood, especially the role of the elusive SRC triplets.

In this work we study the asymptotic form of the nuclear wave-function using the Coupled Cluster (CC) expansion [51, 52], aiming to put the GCF on a more solid theoretical grounds. In addition, the CC expansion provides a systematic way to include higher order corrections, e.g. 3-body SRCs, beyond the leading 2-body SRC term of the asymptotic expansion of the many-body wave-function. Here, we limit our attention to Hamiltonians containing only 2-body interaction, postponing the discussion of 3-body forces to future works.

The paper is organized as follows. In section II we provide a short introduction to the CC expansion method. Then, in section III we discuss the momentum basis and its merits. The derivation of the high-momentum asymptotic equations governing the behavior of two-body and three-body SRCs is presented in section IV. In section V we focus on two-body correlations and analyze their universal behavior. Three-body effects are then analyzed in section VI, where we derive the appropriate universal equation and show its relation to the solution of the zero-energy three-body problem. For the sake of brevity some more technical details are presented in the appendix.

II Coupled cluster theory

The general form of a Hamiltonian describing a many-particle system interacting via two-body potential V^\hat{V} is given by

H^H^0+U^+V^\displaystyle\hat{H}\equiv\hat{H}_{0}+\hat{U}+\hat{V} (4)
=rϵr𝒓𝒓+rrUrr𝒓𝒓+14rsrsVrsrs𝒓𝒔𝒔𝒓,\displaystyle=\sum_{r}\epsilon_{r}\boldsymbol{r}^{\dagger}\boldsymbol{r}+\sum_{rr^{\prime}}U^{r}_{r^{\prime}}{\boldsymbol{r}}^{\dagger}{\boldsymbol{r}}^{\prime}+\frac{1}{4}\sum_{rsr^{\prime}s^{\prime}}V^{rs}_{r^{\prime}s^{\prime}}\boldsymbol{r}^{\dagger}\boldsymbol{s}^{\dagger}{\boldsymbol{s}}^{\prime}{\boldsymbol{r}}^{\prime}, (5)

where H^0\hat{H}_{0} is the “zero-order” or unperturbed Hamiltonian (not necessarily the free Hamiltonian), and U^\hat{U} is the residual one-body interaction. The operators 𝒓,𝒔,{\boldsymbol{r}},{\boldsymbol{s}},\ldots are the usual fermionic ladder operators corresponding to the single particle eigenstates |r,|s,\ket{r},\ket{s},\ldots of H^0\hat{H}_{0}, i.e. H^0|r=ϵr|r\hat{H}_{0}\ket{{r}}=\epsilon_{r}\ket{{r}}, or equivalently

[H^0,𝒓]=ϵr𝒓[H^0,𝒓]=ϵr𝒓,[\hat{H}_{0},{\boldsymbol{r}}^{\dagger}]=\epsilon_{r}{\boldsymbol{r}}^{\dagger}\qquad[\hat{H}_{0},{\boldsymbol{r}}]=-\epsilon_{r}{\boldsymbol{r}}, (6)

where [A^,B^][\hat{A},\hat{B}] is the regular commutator. They obey the anti-commutation relations

{𝒓,𝒔}=0,{𝒓,𝒔}=0,{𝒓,𝒔}=δrs.\{{\boldsymbol{r}},{\boldsymbol{s}}\}=0,\hskip 20.00003pt\{{\boldsymbol{r}}^{\dagger},{\boldsymbol{s}}^{\dagger}\}=0,\hskip 20.00003pt\{{\boldsymbol{r}}^{\dagger},{\boldsymbol{s}}\}=\delta_{rs}. (7)

In the following we will use the notation |r1r2rA\ket{r_{1}r_{2}\ldots r_{A}} to denote normalized antisymmetrized AA-body states and |r1r2rA)|r_{1}r_{2}\ldots r_{A}) to denote the simple, non-symmetrized, many-body states, e.g. |rs=12[|rs)|sr)]\ket{rs}=\frac{1}{\sqrt{2}}[|rs)-|sr)]. The matrix elements of the 2-body potential V^\hat{V} are then given by

Vrsrs=rs|V^|rs=(rs|V^|rs)(rs|V^|sr).V^{rs}_{r^{\prime}s^{\prime}}=\bra{rs}\hat{V}\ket{r^{\prime}s^{\prime}}=(rs|\hat{V}|r^{\prime}s^{\prime})-(rs|\hat{V}|s^{\prime}r^{\prime}). (8)

The starting point of the CC method is a reference Slater-determinant state |Φ0\ket{\Phi_{0}}, composed of AA single particle states. In general, a wave function |Ψ|\Psi\rangle is a linear combination of all such Slater determinants. These determinants can be organized in a systematic way, by considering first the determinants obtained replacing a state occupied in |Φ0\ket{\Phi_{0}} with a state not occupied in |Φ0\ket{\Phi_{0}}, than replacing two such states, and so on. Following the convention of Shavitt & Bartlett [53], we use the letters i,j,,ni,j,\ldots,n to denote ”hole” states, i.e. single-particle states that are occupied in |Φ0|\Phi_{0}\rangle, and the letters a,b,,fa,b,\ldots,f to denote ”particle” states, i.e. single-particle states that are not occupied in |Φ0|\Phi_{0}\rangle. r,s,,wr,s,\ldots,w will be used to denote both states. Therefore,

𝒊|Φ0=0,and𝒂|Φ0=0.\displaystyle\boldsymbol{i}^{\dagger}\ket{\Phi_{0}}=0\;,\quad\text{and}\quad\boldsymbol{a}\ket{\Phi_{0}}=0.{} (9)

The interacting many-body state |Ψ|\Psi\rangle, an eigenstate of H^\hat{H}, is written in the CC formulation as

|Ψ=eT^|Φ0,whereT^=nT^n,\ket{\Psi}=e^{\hat{T}}\ket{\Phi_{0}},\quad\text{where}\quad\hat{T}=\sum_{n}\hat{T}_{{n}}, (10)

and

T^n=1n!2a1an,i1inti1i2ina1a2an𝒂1𝒂2𝒊2𝒊1\hat{T}_{{n}}=\frac{1}{n!^{2}}\sum_{a_{1}\ldots a_{n},i_{1}\ldots i_{n}}t^{a_{1}a_{2}\ldots a_{n}}_{i_{1}i_{2}\ldots i_{n}}\boldsymbol{a}_{1}^{\dagger}\boldsymbol{a}_{2}^{\dagger}\cdots\boldsymbol{i}_{2}\boldsymbol{i}_{1} (11)

is the nn-particle, nn-hole (npnhnpnh) cluster operator.

To determine the amplitudes ti1i2ina1a2ant^{a_{1}a_{2}\ldots a_{n}}_{i_{1}i_{2}\ldots i_{n}}, a set of non-linear equations, the CC equations, can be obtained by projecting the Schrödinger equation on an npnhnpnh state |Φijab𝒂𝒃𝒋𝒊|Φ0\ket{\Phi^{ab\cdots}_{ij\cdots}}\equiv{\boldsymbol{a}}^{\dagger}{\boldsymbol{b}}^{\dagger}\cdots{\boldsymbol{j}}{\boldsymbol{i}}\ket{\Phi_{0}}. The full derivation of the CC equations is given, e.g., in Ref. [53]. Omitting the 1-body potential term U^\hat{U} and the 1p1h1p1h cluster operator T^1\hat{T}_{{1}}, the two- and three-body CC equations are given by

0=Φijab|V^\displaystyle 0=\bra{\Phi_{ij}^{ab}}\hat{V} +[H^0,T^2]+[V^,T^2]+12[[V^,T^2],T^2]\displaystyle+[\hat{H}_{0},\hat{T}_{{2}}]+[\hat{V},\hat{T}_{{2}}]+\frac{1}{2}[[\hat{V},\hat{T}_{{2}}],\hat{T}_{{2}}] (12)
+[V^,T^3]+[V^,T^4]|Φ0,\displaystyle+[\hat{V},\hat{T}_{{3}}]+[\hat{V},\hat{T}_{{4}}]\ket{\Phi_{0}}, (13)
0=Φijkabc|\displaystyle 0=\bra{\Phi_{ijk}^{abc}} [H^0,T^3]+[V^,T^2]+12[[V^,T^2],T^2]+[V^,T^3]\displaystyle[\hat{H}_{0},\hat{T}_{{3}}]+[\hat{V},\hat{T}_{{2}}]+\frac{1}{2}[[\hat{V},\hat{T}_{{2}}],\hat{T}_{{2}}]+[\hat{V},\hat{T}_{{3}}] (14)
+[[V^,T^2],T^3]+[V^,T^4]+[V^,T^5]|Φ0.\displaystyle+[[\hat{V},\hat{T}_{{2}}],\hat{T}_{{3}}]+[\hat{V},\hat{T}_{{4}}]+[\hat{V},\hat{T}_{{5}}]\ket{\Phi_{0}}. (15)

III Momentum basis states

To study SRCs it is most convenient to work with single-particle basis states, i.e. the eigenstate of H^0\hat{H}_{0}, that have well defined momentum. This choice is natural for an infinite system, like nuclear matter - see e.g. [54, 55], but it might seem rather odd for describing a bound nucleus which is a compact object. However, large nuclei have relatively constant density and far from the surface behave like an infinite nuclear system. Thus, we set the problem in a box of size LL with periodic boundary conditions. For LL larger than the nucleus size, the wave-function and the binding energy approach very fast the free space (LL\to\infty) values and we need not worry about the impact of the boundary conditions on the nuclear SRCs.

Assuming 𝒑=(p1,p2,p3){\boldsymbol{p}}=(p_{1},p_{2},p_{3}) to be a triad of integers, the basis states {|𝒑}\{\ket{{\boldsymbol{p}}}\}

𝒙|𝒑=1Ωei2πL𝒙𝒑,𝒑|𝒑=δ𝒑,𝒑,\bra{{\boldsymbol{x}}}\ket{{\boldsymbol{p}}}=\frac{1}{\sqrt{\Omega}}e^{i\frac{2\pi}{L}{\boldsymbol{x}}\cdot{\boldsymbol{p}}},\quad\bra{{\boldsymbol{p}}}\ket{{\boldsymbol{p}}^{\prime}}=\delta_{{\boldsymbol{p}},{\boldsymbol{p}}^{\prime}}, (16)

with Ω=L3\Omega=L^{3}, is a complete set of orthonormal states, which combined with the spin and isospin degrees of freedom form our single-particle basis. A natural choice for |Φ0\ket{\Phi_{0}}, the starting point of the CC expansion, is a Slater determinant composed of the AA lowest kinetic energy single-particle states.

If there is a well defined Fermi momentum pFp_{F}, such that all the hole states are momentum states with momentum smaller than pFp_{F}, while particle states have momentum larger than pFp_{F}, then the system is called a closed shell system. To simplify matters, in the following we shall restrict our the discussion to such closed shell systems only.

Working with this single-particle momentum basis, H^0\hat{H}_{0} coincides with the kinetic energy operator and therefore U^=0\hat{U}=0. The Slater-determinant |Φ0\ket{\Phi_{0}}, as well as the npnhnpnh states |Φijab\ket{\Phi^{ab\cdots}_{ij\cdots}}, are a product of single-particle momentum states, hence they are eigenstates of the total center of mass (CM) momentum operator 𝑷^CM\hat{{\boldsymbol{P}}}_{CM}. The two-body potential is translational invariant, hence the CM momentum is a good quantum number, and the wave-function |Ψ\ket{\Psi} is also an eigenstate of 𝑷^CM\hat{{\boldsymbol{P}}}_{CM},

𝑷^CM|Ψ=𝑷^CMeT^|Φ0=𝑷CM|Ψ.\displaystyle\hat{{\boldsymbol{P}}}_{CM}\ket{\Psi}=\hat{{\boldsymbol{P}}}_{CM}e^{\hat{T}}\ket{\Phi_{0}}={\boldsymbol{P}}_{CM}\ket{\Psi}.{} (17)

Closing the last equation with Φ0|\bra{\Phi_{0}} and acting with 𝑷^CM\hat{{\boldsymbol{P}}}_{CM} once to the left and once to the right, and noting that Φ0|Ψ0\bra{\Phi_{0}}\ket{\Psi}\neq 0, we must conclude that |Φ0\ket{\Phi_{0}} and |Ψ\ket{\Psi} share the same eigenvalue of the total momentum 𝑷CM{\boldsymbol{P}}_{CM}.

We may now repeat the same argument for the 1p1h1p1h states. Closing Eq. (17) with Φia|\bra{\Phi_{i}^{a}} and using Φia|Ψ=tia\langle\Phi_{i}^{a}|\Psi\rangle=t_{i}^{a} we get

(𝒑a𝒑i)tia=𝟎,({\boldsymbol{p}}_{a}-{\boldsymbol{p}}_{i})t^{a}_{i}=\boldsymbol{0}, (18)

which for all closed shell systems implies [56, 54]

tia=0,t^{a}_{i}=0, (19)

because (𝒑a𝒑i)𝟎({\boldsymbol{p}}_{a}-{\boldsymbol{p}}_{i})\neq\boldsymbol{0}, as 𝒑a{\boldsymbol{p}}_{a} corresponds to a particle state while 𝒑i{\boldsymbol{p}}_{i} to a hole state. Thus, with this choice of basis states, T^1\hat{T}_{{1}} is eliminated from the CC expansion, as was assumed in Eqs. (12) and (14).

Considering now the 2p2h2p2h states, multiplying Eq. (17) by Φijab|\bra{\Phi^{ab}_{ij}} one gets [55]

(𝒑a+𝒑b𝒑i𝒑j)tijab=𝟎.({\boldsymbol{p}}_{a}+{\boldsymbol{p}}_{b}-{\boldsymbol{p}}_{i}-{\boldsymbol{p}}_{j})t^{ab}_{ij}=\boldsymbol{0}. (20)

This implies that T^2\hat{T}_{{2}} conserves momentum, i.e. tijab=0t_{ij}^{ab}=0 if 𝒑a+𝒑b𝒑i𝒑j𝟎{\boldsymbol{p}}_{a}+{\boldsymbol{p}}_{b}-{\boldsymbol{p}}_{i}-{\boldsymbol{p}}_{j}\neq\boldsymbol{0}. It can be similarly shown that for a closed shell system all amplitude operators T^n\hat{T}_{{n}} must conserve momentum.

IV Coupled cluster amplitudes in the high momentum limit

SRCs are associated with high momentum particles. To understand their role in the many-body wave-function we need to study the high momentum behavior of the CC amplitudes T^n\hat{T}_{{n}} as dictated by Eqs. (12) and (14). In the following we will assume a,ba,b and cc to be highly excited states corresponding to momenta pa,pb,pcpFp_{a},p_{b},p_{c}\gg p_{\text{F}}. We note that in order for the wave-function to be properly normalized the CC amplitudes T^n\hat{T}_{{n}} must vanish in this limit, e.g. tijkabc0t^{abc}_{ijk}\to 0 when a,b,ca,b,c\to\infty.

For a system of fermions, we expect the CC amplitudes to admit the natural hierarchy, where double excitations are much more significant than three-body excitations which on their part are more important than the four-body excitations, etc. It follows that the contributions of [V^,T^3][\hat{V},\hat{T}_{{3}}] and [V^,T^4][\hat{V},\hat{T}_{{4}}] to the 2-body equation can be neglected. Similarily, the terms [V^,T^4][\hat{V},\hat{T}_{{4}}] and [V^,T^5][\hat{V},\hat{T}_{{5}}] can be neglected in the 3-body CC equation.

In order to understand the behaviour of the CC amplitudes in the high momentum limit, let us inspect the T^2\hat{T}_{{2}} equation, Eq. (12), in the limit pa,pbp_{a},p_{b}\to\infty. In this case, the leading terms are the source term VijabV^{ab}_{ij} and the kinetic energy term [H^0,T^2][\hat{H}_{0},\hat{T}_{{2}}]. Retaining only these terms leads to the well known asymptotic result

tijab1EijabVijab,t^{ab}_{ij}\rightarrow-\frac{1}{E^{ab}_{ij}}V^{ab}_{ij}, (21)

were EijabE^{ab}_{ij} is the excitation energy given by the relation

Ei1i2ina1a2an(ϵa1+ϵa2+ϵan)(ϵi1++ϵin).E^{a_{1}a_{2}\ldots a_{n}}_{i_{1}i_{2}\ldots i_{n}}\equiv(\epsilon_{a_{1}}+\epsilon_{a_{2}}+\ldots\epsilon_{a_{n}})-(\epsilon_{i_{1}}+\ldots+\epsilon_{i_{n}}). (22)

If, for pa,pbp_{a},p_{b}\to\infty, the potential matrix elements VijabV^{ab}_{ij} are independent of the exact holes states, i.e. VijabV0i0jabV^{ab}_{ij}\approx V^{ab}_{0_{i}0_{j}}, with 0i0_{i} being a zero-momentum state (used loosely to indicate the lowest momentum state with the same quantum numbers as the state ii), the asymptotic two-body amplitude presented in Eq. (21) is universal in the limited sense. That is, tijab1E00abV00abt^{ab}_{ij}\approx-\frac{1}{E^{ab}_{00}}V^{ab}_{00} is independent of the number of nucleons AA and the specifics of the nuclear state. On the other hand, it depends on the potential - therefore its universality is limited. This form of asymptotic behaviour was first suggested by Amado [57] exploring the asymptotic form of the nuclear momentum distribution. It turns out, however, that although Eq. (21) is asymptotically correct, it is valid only at extremely high momentum, larger than 10 fm-1, making it impractical for actual calculations [58]. Consequently, in order to get a reasonable description of the asymptotic nuclear wave-function, we must retain more terms besides the source term and the [H^0,T^n][\hat{H}_{0},\hat{T}_{{n}}] terms in the CC equations.

With the 3, and 4-body amplitudes neglected, the CC T^2\hat{T}_{{2}} equation, Eq. (12), takes the form

0=Φijab|V^+[H^0,T^2]+[V^,T^2]+12[[V^,T^2],T^2]|Φ0.0=\bra{\Phi_{ij}^{ab}}\hat{V}+[\hat{H}_{0},\hat{T}_{{2}}]+[\hat{V},\hat{T}_{{2}}]+\frac{1}{2}[[\hat{V},\hat{T}_{{2}}],\hat{T}_{{2}}]\ket{\Phi_{0}}. (23)

Comparing now the terms [V^,T^2][\hat{V},\hat{T}_{{2}}], and [[V^,T^2],T^2][[\hat{V},\hat{T}_{{2}}],\hat{T}_{{2}}] we note that for the latter we get the following matrix elements, ignoring combinatorical factors,

Vdekltikabtjlde,Vdekltklabtijde,Vdekltikadtjlbe,Vdekltijadtklbe.V^{kl}_{de}t^{ab}_{ik}t^{de}_{jl},\quad V^{kl}_{de}t^{ab}_{kl}t^{de}_{ij},\quad V^{kl}_{de}t^{ad}_{ik}t^{be}_{jl},\quad V^{kl}_{de}t^{ad}_{ij}t^{be}_{kl}. (24)

Here, for brevity, we use the Einstein convention assuming implicit summation on repeated lower and upper indices. In the high momentum pa,pbp_{a},p_{b}\to\infty and low density i,j,k,l0i,j,k,l\to 0 limit, these terms take the form

2Vde00t00det00ab,and2Vde00t00adt00be.2V^{00}_{de}t^{de}_{00}t^{ab}_{00},\quad\text{and}\quad 2V^{00}_{de}t^{ad}_{00}t^{be}_{00}. (25)

The first of these terms is nothing but an energy shift, a correction to the excitation energy EijabE^{ab}_{ij}, which we can neglect in the high momentum limit. We note that the second term is zero unless 𝒑d=𝒑a{\boldsymbol{p}}_{d}=-{\boldsymbol{p}}_{a}, 𝒑e=𝒑b{\boldsymbol{p}}_{e}=-{\boldsymbol{p}}_{b} and 𝒑b=𝒑a{\boldsymbol{p}}_{b}=-{\boldsymbol{p}}_{a}. This term is clearly suppressed by a factor of t00abt^{ab}_{00} with respect to [V^,T^2][\hat{V},\hat{T}_{{2}}], and thus can be neglected as well.

Considering now the T^3\hat{T}_{{3}} equation, Eq. (14). After neglecting the T^4\hat{T}_{{4}}, T^5\hat{T}_{{5}} as well as the [[V^,T^2],T^3][[V^,T^2],T^2][[\hat{V},\hat{T}_{{2}}],\hat{T}_{{3}}]\ll[[\hat{V},\hat{T}_{{2}}],\hat{T}_{{2}}] terms, we remain with

0=Φijkabc|\displaystyle 0=\bra{\Phi_{ijk}^{abc}} [H^0,T^3]+[V^,T^2]\displaystyle[\hat{H}_{0},\hat{T}_{{3}}]+[\hat{V},\hat{T}_{{2}}]
+12[[V^,T^2],T^2]+[V^,T^3]|Φ0.\displaystyle+\frac{1}{2}[[\hat{V},\hat{T}_{{2}}],\hat{T}_{{2}}]+[\hat{V},\hat{T}_{{3}}]\ket{\Phi_{0}}.{} (26)

Comparing again the [V^,T^2][\hat{V},\hat{T}_{{2}}], and [[V^,T^2],T^2][[\hat{V},\hat{T}_{{2}}],\hat{T}_{{2}}], we see that the only terms that survive in the high momentum/low density limit are respectively Ve0abt00ceV^{ab}_{e0}t^{ce}_{00}, and Vefa0t00bet00cfV^{a0}_{ef}t^{be}_{00}t^{cf}_{00}. Thus as before, the double commutator term is suppressed by a factor of t00abt^{ab}_{00} and can be neglected.

Summing up, in the limit of high momenta, we expect the two- and three-body CC amplitudes to obey the equations:

0\displaystyle 0 =Φijab|[H^0,T^2]+[V^,T^2]+V^|Φ0,\displaystyle=\bra{\Phi_{ij}^{ab}}[\hat{H}_{0},\hat{T}_{{2}}]+[\hat{V},\hat{T}_{{2}}]+\hat{V}\ket{\Phi_{0}}, (27)
0\displaystyle 0 =Φijkabc|[H^0,T^3]+[V^,T^3]+[V^,T^2]|Φ0.\displaystyle=\bra{\Phi_{ijk}^{abc}}[\hat{H}_{0},\hat{T}_{{3}}]+[\hat{V},\hat{T}_{{3}}]+[\hat{V},\hat{T}_{{2}}]\ket{\Phi_{0}}. (28)

In the following sections we will analyze these equations.

V The 2-body amplitude

In section IV we argued that asymptotically the 2-body CC equation takes the form of Eq. (27). In order to evaluate this equation, we note that there can be no contractions between the operators 𝒂𝒃𝒋𝒊\boldsymbol{a}^{\dagger}\boldsymbol{b}^{\dagger}\boldsymbol{j}\boldsymbol{i} that appear in the bra state, hence all the labels a,b,ia,b,i and jj has to appear on the amplitude and potential operators. Therefore Φijab|V^|Φ0=Vijab\bra{\Phi_{ij}^{ab}}\hat{V}\ket{\Phi_{0}}=V^{ab}_{ij}, and Φijab|[H^0,T^2]|Φ0=Eijabtijab\bra{\Phi_{ij}^{ab}}\left[{\hat{H}_{0},\hat{T}_{{2}}}\right]\ket{\Phi_{0}}=E^{ab}_{ij}t^{ab}_{ij}. To evaluate the commutator [V^,T^2]=V^T^2T^2V^\left[{\hat{V},\hat{T}_{{2}}}\right]=\hat{V}\hat{T}_{{2}}-\hat{T}_{{2}}\hat{V}, we note that all the operators of V^\hat{V} in V^T^2\hat{V}\hat{T}_{{2}} can be moved to the right of T^2\hat{T}_{{2}} and then the term will cancel with T^2V^\hat{T}_{{2}}\hat{V}. In the process, all possible contractions between V^\hat{V} and T^2\hat{T}_{{2}} will arise, i.e. at least one contraction should be made between them. This commutators yield 5 distinct terms, that combined with the potential and the H^0\hat{H}_{0} terms results in the linear Coupled-Cluster doublets (CCD) equation

0\displaystyle 0 =Vijab+Eijabtijab+12Vdeabtijde\displaystyle=V^{ab}_{ij}+E^{ab}_{ij}t^{ab}_{ij}+\frac{1}{2}V^{ab}_{de}t^{de}_{ij}
+12Vijkltklab+Vidaktjkbd+Vkdaktijbd+Viklktjlab\displaystyle+\frac{1}{2}V^{kl}_{ij}t^{ab}_{kl}+V^{ak}_{id}t^{bd}_{jk}+V^{ak}_{kd}t^{bd}_{ij}+V^{lk}_{ik}t^{ab}_{jl}
+permutations.\displaystyle+\text{permutations}\;.{} (29)

The title ‘permutations’ stands for anti-symmetrization with respect to the indices abab or ijij when not placed on the same matrix elements. The summation VkwkvV^{kv}_{kw} is performed only on hole states as the string 𝒓𝒔\boldsymbol{r}^{\dagger}\boldsymbol{s} of V^\hat{V}, where both 𝒓,𝒔{\boldsymbol{r}},{\boldsymbol{s}} are particle operators, is already normal ordered and therefore its contraction is zero.

In the limit of high momentum/low density the 2nd line of (V) take either the form V0000t00abV^{00}_{00}t^{ab}_{00} or Va0a0t00abV^{a0}_{a0}t^{ab}_{00}. In both cases these terms enter as small corrections to the excitation energy. It follows that these terms can be neglected for large pa,pbp_{a},p_{b} with respect to the terms appearing on the first line.

Refining this argument, due to momentum conservation we expect that in the limit pa,pbp_{a},p_{b}\to\infty all the T^2\hat{T}_{{2}} terms on the 2nd and 3rd lines of Eq. (V), will be either exactly or at least approximately equal to tijabt^{ab}_{ij} since the hole states carry only low momentum of the order of pFp_{\text{F}} and we assume weak momenta dependance on the hole states quantum numbers. For example, in VidaktjkbdV^{ak}_{id}t^{bd}_{jk} the momentum 𝒑d{\boldsymbol{p}}_{d} associated with the state dd must be of the order 𝒑d=𝒑a+(pF){\boldsymbol{p}}_{d}={\boldsymbol{p}}_{a}+\order{p_{\text{F}}} implying that tjkbdtijbat^{bd}_{jk}\approx t^{ba}_{ij}. Comparing these terms to the term EijabtijabE^{ab}_{ij}t^{ab}_{ij}, we see that for large enough excitations

EijabklVijkl,kdVidak,kdVkdak,klViklkE^{ab}_{ij}\gg\sum_{kl}V^{kl}_{ij},\;\sum_{kd}V^{ak}_{id},\;\sum_{kd}V^{ak}_{kd},\;\sum_{kl}V^{lk}_{ik} (30)

and these terms can be neglected. It is important to observe that the neglected terms are all intensive and do not scale with the size of the system.

The resulting 2-body amplitude equation is then

0\displaystyle 0 =tijab+1EijabVijab+12EijabVdeabtijde,\displaystyle=t^{ab}_{ij}+\frac{1}{E^{ab}_{ij}}V^{ab}_{ij}+\frac{1}{2E^{ab}_{ij}}V^{ab}_{de}t^{de}_{ij}\;,{} (31)

which is nothing but particle-particle ladder approximation of the CCD equation, applied for example in Ref. [54] to estimate the nuclear matter equation of state. In the following we will use the notation T^2\hat{T}^{\infty}_{{2}} to denote the solution of Eq. (31) in the non-symmetrized basis with Eijϵi+ϵj0E^{ij}\equiv\epsilon_{i}+\epsilon_{j}\to 0. As it is a linear equation, T^2\hat{T}^{\infty}_{{2}} is unique. We will show in appendix A that indeed 𝒂𝒃|T^2|𝒊𝒋𝒂𝒃|T^2|𝒊𝒋\bra{\boldsymbol{ab}}\hat{T}_{{2}}\ket{\boldsymbol{ij}}\to\bra{\boldsymbol{ab}}\hat{T}^{\infty}_{{2}}\ket{\boldsymbol{ij}} as pa,pbp_{a},p_{b}\to\infty.

We can now discuss the properties of T^2\hat{T}^{\infty}_{{2}}. As stated above, the cluster operator T^2\hat{T}^{\infty}_{{2}} is defined as the solution of the equation

0\displaystyle 0 =(t)ijab+1EabVijab+12EabVdeab(t)ijde,\displaystyle=\left({t^{\infty}}\right)^{ab}_{ij}+\frac{1}{E^{ab}}V^{ab}_{ij}+\frac{1}{2E^{ab}}V^{ab}_{de}\left({t^{\infty}}\right)^{de}_{ij},{} (32)

where Eab=ϵa+ϵbE^{ab}=\epsilon_{a}+\epsilon_{b}. A similar result was derived by Zabolitzky in Ref. [58]. To analyze this equation, it is convenient to introduce particle-particle and the hole-hole projection operators,

Q2=de|de)(de|,P2=lm|lm)(lm|,{Q}_{{2}}=\sum_{de}|{d}{e})({d}{e}|,\quad{P}_{{2}}=\sum_{lm}|{l}{m})({l}{m}|, (33)

and the Green’s function

G^0(E)=1EH^0+iε.\hat{G}_{0}(E)=\frac{1}{E-\hat{H}_{0}+i\varepsilon}. (34)

Eq. (32) can then be written as

T^2=Q2G^0(0)V^T^2+Q2G^0(0)V^P2,\hat{T}^{\infty}_{{2}}={Q}_{{2}}\hat{G}_{0}(0)\hat{V}\hat{T}^{\infty}_{{2}}+{Q}_{{2}}\hat{G}_{0}(0)\hat{V}{P}_{{2}}, (35)

and formally solved to yield

T^2=11Q2G^0(0)V^Q2G^0(0)V^P2.\hat{T}^{\infty}_{{2}}=\frac{1}{1-{Q}_{{2}}\hat{G}_{0}(0)\hat{V}}{Q}_{{2}}\hat{G}_{0}(0)\hat{V}{P}_{{2}}\;. (36)

Clearly, P2T^2=T^2Q2=0{P}_{{2}}\hat{T}^{\infty}_{{2}}=\hat{T}^{\infty}_{{2}}{Q}_{{2}}=0 as expected from a cluster operator. Using the relation Q2H^0P2=0{Q}_{{2}}\hat{H}_{0}{P}_{{2}}=0 the solution (36) can be rewritten as (see appendix B)

T^2\displaystyle\hat{T}^{\infty}_{{2}} =1Q2(0+iεH^)Q2Q2H^P2.\displaystyle=\frac{1}{{Q}_{{2}}(0+i\varepsilon-\hat{H}){Q}_{{2}}}{Q}_{{2}}\hat{H}{P}_{{2}}\;.{} (37)

Before proceeding, we note that P2{P}_{{2}} is not eqivalent to Q¯2=1Q2\bar{Q}_{{2}}=1-{Q}_{{2}} the complement of Q2{Q}_{{2}}, as Q¯2\bar{Q}_{{2}} must include not only hole-hole states but also particle-hole states. For infinite nuclear matter we expect however, that T^2\hat{T}^{\infty}_{{2}} is translational invariant and therefore we can consider only pairs with zero CM momentum. For such pairs, there are no particle-hole contributions and we can replace P2{P}_{{2}} by Q¯2\bar{Q}_{{2}}. In this subspace

T^2=1Q2(0+iεH^)Q2Q2H^Q¯2.\hat{T}^{\infty}_{{2}}=\frac{1}{{Q}_{{2}}(0+i\varepsilon-\hat{H}){Q}_{{2}}}{Q}_{{2}}\hat{H}\bar{Q}_{{2}}\;. (38)

Comparing now Eq. (38) with the Bloch-Horowitz equations [59],

Q¯2|Ψ\displaystyle\bar{Q}_{{2}}|\Psi\rangle =1Q¯2(E+iεH^)Q¯2Q¯2H^Q2|Ψ\displaystyle=\frac{1}{\bar{Q}_{{2}}(E+i\varepsilon-\hat{H})\bar{Q}_{{2}}}\bar{Q}_{{2}}\hat{H}{Q}_{{2}}|\Psi\rangle (39)
Q2|Ψ\displaystyle{Q}_{{2}}|\Psi\rangle =1Q2(E+iεH^)Q2Q2H^Q¯2|Ψ,\displaystyle=\frac{1}{{Q}_{{2}}(E+i\varepsilon-\hat{H}){Q}_{{2}}}{Q}_{{2}}\hat{H}\bar{Q}_{{2}}|\Psi\rangle\;, (40)

it is clear that T^2\hat{T}^{\infty}_{{2}} is nothing but the zero-energy 2-body Bloch-Horowitz operator

O^2B.H.=1Q2(0+iεH^)Q2Q2H^Q¯2.\hat{O}_{2}^{\text{B.H.}}=\frac{1}{{Q}_{{2}}(0+i\varepsilon-\hat{H}){Q}_{{2}}}{Q}_{{2}}\hat{H}\bar{Q}_{{2}}. (41)

This operator fulfills the relation O^2B.H.|Ψ2=Q2|Ψ2\hat{O}_{2}^{\text{B.H.}}\ket{\Psi_{2}}={Q}_{{2}}\ket{\Psi_{2}} for any zero energy eigenstate |Ψ2\ket{\Psi_{2}} that obeys H^|Ψ2=0\hat{H}\ket{\Psi_{2}}=0. It follows that if H^|Ψ2=0\hat{H}\ket{\Psi_{2}}=0 and 𝑷^CM|Ψ2=0\hat{\boldsymbol{P}}_{CM}\ket{\Psi_{2}}=0, then

Q2|Ψ2=T^2|Ψ2.{Q}_{{2}}\ket{\Psi_{2}}=\hat{T}^{\infty}_{{2}}\ket{\Psi_{2}}. (42)

Inspecting eqs. (37) and (42), we can conclude that (i) The asymptotic 2-body behavior of T^2\hat{T}_{{2}}, and therefore of the many-body wave-function, is related to the zero-energy solutions of the 2-body problem. (ii) The relation to the zero-energy solutions show the universality of the asymptotic behavior in the limited sense, that it does not depend on the system, but depends on the potential.

VI The 3-body amplitude

As we have argued in Sec. IV, the behavior of the 3-body amplitude T^3\hat{T}_{{3}} at high momentum is dictated by Eq. (28). Explicitly, this equation takes the form

0\displaystyle 0 =EijkabctijkabcVijlatklbcVidabtjkcd\displaystyle=E^{abc}_{ijk}t^{abc}_{ijk}-V^{la}_{ij}t^{bc}_{kl}-V^{ab}_{id}t^{cd}_{jk}
+12Vdeabtijkcde+12Vijlmtklmabc+Vdlaltijkbcd+Vidaltjklbcd+Villmtjkmabc\displaystyle+\frac{1}{2}V^{ab}_{de}t^{cde}_{ijk}+\frac{1}{2}V^{lm}_{ij}t^{abc}_{klm}+V^{al}_{dl}t^{bcd}_{ijk}+V^{al}_{id}t^{bcd}_{jkl}+V^{lm}_{il}t^{abc}_{jkm}
+permutations.\displaystyle+\text{permutations}\;.{} (43)

Here, the first term on the rhs is due to [H^0,T^3][\hat{H}_{0},\hat{T}_{{3}}], the next two terms come from the [V^,T^2][\hat{V},\hat{T}_{{2}}] commutator, and the next five are due to the [V^,T^3][\hat{V},\hat{T}_{{3}}] commutator. The title ‘permutations’ stands for anti-symmetrization with respect to the indices abcabc or ijkijk when not placed on the same matrix elements. Due to momentum conservation, for very large pap_{a} the potential matrix elements VijlaV^{la}_{ij} must vanish, leaving VidabtjkcdV^{ab}_{id}t^{cd}_{jk} as the only source term. In addition, all terms coming from the [V^,T^3][\hat{V},\hat{T}_{{3}}] commutator, except for the first term in the second line (and its corresponding permutations), are approximately proportional to tijkabct^{abc}_{ijk}. Therefore, for excitation energy EijkabcE^{abc}_{ijk} large enough

EijkabclmVijlm,ldVdlal,ldVidal,mlVillm,E^{abc}_{ijk}\gg\sum_{lm}V^{lm}_{ij},\;\sum_{ld}V^{al}_{dl}\;,\sum_{ld}V^{al}_{id}\;,\sum_{ml}V^{lm}_{il}\;, (44)

and the corresponding terms can be neglected in comparison to the free term EijkabctijkabcE^{abc}_{ijk}t^{abc}_{ijk}. As a consequence only the terms 12Vdeabtijkcde\frac{1}{2}V^{ab}_{de}t^{cde}_{ijk} remain. Utilizing these observations, and defining the symmetrization operator 𝒮^1231+(123)+(132){\cal\hat{S}}_{123}\equiv 1+(123)+(132) where (123)(123) is the permutation operator, equation (VI) takes the form

0=tijkabc\displaystyle 0=t^{abc}_{ijk} +𝒮^abc(𝒮^ijk(Vidabtjkdc))Eijkabc+𝒮^abc(Vdeabtijkcde)2Eijkabc.\displaystyle+\frac{{\cal\hat{S}}_{abc}({\cal\hat{S}}_{ijk}(V^{ab}_{id}t^{dc}_{jk}))}{E^{abc}_{ijk}}+\frac{{\cal\hat{S}}_{abc}(V^{ab}_{de}t^{cde}_{ijk})}{2E^{abc}_{ijk}}\;.{} (45)

As in the 2-body case we define T^3\hat{T}^{\infty}_{{3}} to be the solution of eq. (45) in the limit EijkabcEabcE^{abc}_{ijk}\to E^{abc}, and tjkdc(t)jkdct^{dc}_{jk}\to\left({t^{\infty}}\right)^{dc}_{jk}. We show in appendix C that 𝒂𝒃𝒄|T^3|𝒊𝒋𝒌𝒂𝒃𝒄|T^3|𝒊𝒋𝒌\bra{\boldsymbol{abc}}\hat{T}_{{3}}\ket{\boldsymbol{ijk}}\to\bra{\boldsymbol{abc}}\hat{T}^{\infty}_{{3}}\ket{\boldsymbol{ijk}} as pa,pb,pcp_{a},p_{b},p_{c}\to\infty.

To analyze T^3\hat{T}^{\infty}_{{3}} we write eq. (45) in first quantization using the non-symmetrized basis defined above. In the 3-body case, the relation between the anti-symmetrized matrix elements and the non symmetrized ones is

rst|\displaystyle\bra{rst} O^|uvw\displaystyle\hat{O}\ket{uvw}
=𝒮^uvw[(rst|O^|uvw)(rst|O^|vuw)],\displaystyle={\cal\hat{S}}_{uvw}\left[(rst|\hat{O}|uvw)-(rst|\hat{O}|vuw)\right],{} (46)

and for a 2-body operator closed by 3-particle states

(rst|O^2\displaystyle(rst|\hat{O}_{2} |uvw)i=13(rst|O^2(i)|uvw)\displaystyle|uvw)\equiv\sum_{i=1}^{3}(rst|\hat{O}_{2}(i)|uvw){} (47)

where O^2(i)\hat{O}_{2}(i) does not act on the ii’th particles, e.g. (rst|O^2(3)|uvw)=(rs|O^2|uv)δt,w(rst|\hat{O}_{2}(3)|uvw)=(rs|{\hat{O}_{2}}|uv)\delta_{t,w}. With the projection operators

Q3=def|def)(def|,P3=lmn|lmn)(lmn|,{Q}_{{3}}=\sum_{def}|def)(def|,\quad{P}_{{3}}=\sum_{lmn}|lmn)(lmn|, (48)

the asymptotic equation for T^3\hat{T}^{\infty}_{{3}} can be written as

T^3\displaystyle\hat{T}^{\infty}_{{3}} =Q3G^0(0)V^T^3+Q3G^0(0)V^T^2P3\displaystyle={Q}_{{3}}\hat{G}_{0}\left({0}\right)\hat{V}\hat{T}^{\infty}_{{3}}+{Q}_{{3}}\hat{G}_{0}\left({0}\right)\hat{V}\hat{T}^{\infty}_{{2}}{P}_{{3}}
=11Q3G^0(0)V^Q3G^0(0)V^T^2P3\displaystyle=\frac{1}{1-{Q}_{{3}}\hat{G}_{0}\left({0}\right)\hat{V}}{Q}_{{3}}\hat{G}_{0}\left({0}\right)\hat{V}\hat{T}^{\infty}_{{2}}{P}_{{3}}
=1Q3(0+iεH^)Q3Q3H^T^2P3.\displaystyle=\frac{1}{{Q}_{{3}}(0+i\varepsilon-\hat{H}){Q}_{{3}}}{Q}_{{3}}\hat{H}\hat{T}^{\infty}_{{2}}{P}_{{3}}\;.{} (49)

Comparing eq. (VI) with the 3-body Bloch-Horowitz equations [59] and noting that for 2-body interactions Q3HQ¯3=Q3H(Q1P2+Q2P1){Q}_{{3}}H\bar{Q}_{{3}}={Q}_{{3}}H\left({{Q}_{{1}}{P}_{{2}}+{Q}_{{2}}{P}_{{1}}}\right) with Q¯3=1Q3\bar{Q}_{{3}}=1-{Q}_{{3}}

Q¯3|Ψ\displaystyle\bar{Q}_{{3}}|\Psi\rangle =1Q¯3(E+iεH^)Q¯3Q¯3H^Q3|Ψ\displaystyle=\frac{1}{\bar{Q}_{{3}}(E+i\varepsilon-\hat{H})\bar{Q}_{{3}}}\bar{Q}_{{3}}\hat{H}{Q}_{{3}}|\Psi\rangle (50)
Q3|Ψ\displaystyle{Q}_{{3}}|\Psi\rangle =1Q3(E+iεH^)Q3Q3H^Q¯3|Ψ,\displaystyle=\frac{1}{{Q}_{{3}}(E+i\varepsilon-\hat{H}){Q}_{{3}}}{Q}_{{3}}\hat{H}\bar{Q}_{{3}}|\Psi\rangle\;, (51)

we can connect T^3\hat{T}^{\infty}_{{3}} to the zero-energy Bloch-Horowitz operator. Specifically, if |Ψ3|\Psi_{3}\rangle is a zero-energy 3-body eigenstate of H^\hat{H}, and if there is a 3-hole state |α3|\alpha_{3}\rangle such that

T^2|α3\displaystyle\hat{T}^{\infty}_{{2}}|\alpha_{3}\rangle (Q1P2+Q2P1)|Ψ3\displaystyle\approx\left({{Q}_{{1}}{P}_{{2}}+{Q}_{{2}}{P}_{{1}}}\right)|\Psi_{3}\rangle{} (52)

then

T^3|α3Q3|Ψ3\hat{T}^{\infty}_{{3}}\ket{\alpha_{3}}\approx{Q}_{{3}}\ket{\Psi_{3}} (53)

and we can identify the matrix-elements of T^3\hat{T}^{\infty}_{{3}} with the Q3{Q}_{{3}} components of the zero-energy solutions of the Schrödinger equation. In the next section we will argue that eq. (53) approximately holds.

VI.1 The 3-body zero-energy eigenstate

We first note that a zero-energy 3-body eigenstate of the Schrödinger equation, H^|Ψ3=0\hat{H}\ket{\Psi_{3}}=0, can be formally expanded in the CC method as

|Ψ3=𝒩31eT~^|Φ0=𝒩31(1+T~^2+T~^3)|Φ~0,\displaystyle|\Psi_{3}\rangle={\cal N}_{3}^{-1}e^{\hat{\tilde{T}}}|\Phi_{0}\rangle={\cal N}_{3}^{-1}(1+\hat{\tilde{T}}_{{2}}+\hat{\tilde{T}}_{{3}})|\tilde{\Phi}_{0}\rangle,{} (54)

where T~^2,T~^3\hat{\tilde{T}}_{{2}},\hat{\tilde{T}}_{{3}} are the 3-body cluster operators, and

𝒩32=1+Tr(T~^2T~^2)+Tr(T~^3T~^3)\displaystyle{\cal N}_{3}^{2}=1+\Tr\left({\hat{\tilde{T}}_{{2}}^{\dagger}\hat{\tilde{T}}_{{2}}}\right)+\Tr\left({\hat{\tilde{T}}_{{3}}^{\dagger}\hat{\tilde{T}}_{{3}}}\right)\;{} (55)

is a normalization factor. Working with the momentum basis, we note that whereas the AA-body operators T^k\hat{T}_{{k}} are defined with respect to the AA-body Fermi level pFp_{\text{F}}, the 3-body operators T~^2,T~^3\hat{\tilde{T}}_{{2}},\hat{\tilde{T}}_{{3}} are defined with respect to a 3-body reference state which we denote as |000\ket{000}, to indicate that it corresponds to single particle states with momentum which is either zero or very close to zero. We note that in this case the state |000\ket{000} acts as a closed-shell state as the other possible Slater-determinants with zero momenta holes have different conserved quantum numbers, such as of J^z\hat{J}_{z}, and cannot contribute to |Ψ3\ket{\Psi_{3}}. It follows that

|Ψ3\displaystyle|\Psi_{3}\rangle =𝒩31(|000+12lmt~00lm|lm0+12det~00de|de0\displaystyle={\cal N}_{3}^{-1}\left(\ket{000}+\frac{1}{2}\sum_{lm}\tilde{t}^{lm}_{00}\ket{lm0}+\frac{1}{2}\sum_{de}\tilde{t}^{de}_{00}\ket{de0}\right.
+16lmnt~000lmn|lmn+12dlmt~000dlm|dlm\displaystyle\left.+\frac{1}{6}\sum_{lmn}\tilde{t}^{lmn}_{000}\ket{lmn}+\frac{1}{2}\sum_{dlm}\tilde{t}^{dlm}_{000}\ket{dlm}\right.
+12delt~000del|del+16deft~000def|def).\displaystyle\left.+\frac{1}{2}\sum_{del}\tilde{t}^{del}_{000}\ket{del}+\frac{1}{6}\sum_{def}\tilde{t}^{def}_{000}\ket{def}\right)\;.{} (56)

Here, we keep notating the states according to the AA-body Fermi level, e.g. d,e,fd,e,f correspond to particle states while l,m,nl,m,n to hole states. As a result, terms such as |dl0\ket{dl0} cannot appear in the expansion, as momentum conservation implies that if the state dd is above the Fermi level so must be ll.

Before substituting the 3-body wave-function (VI.1) into the QQ-space Bloch-Horowitz equation (51) we note that: (i) The operator Q¯3\bar{Q}_{{3}} kills the 3p0h3p0h states |def\ket{def}. (ii) For 2-body interactions the term Q3H^Q¯3{Q}_{{3}}\hat{H}\bar{Q}_{{3}} annihilates the 0p3h0p3h states, thus

Q3H^Q¯3|Ψ3=\displaystyle{Q}_{{3}}\hat{H}\bar{Q}_{{3}}|\Psi_{3}\rangle= Q3H^Q¯3|Ψ3(1p,2p)\displaystyle{Q}_{{3}}\hat{H}\bar{Q}_{{3}}|\Psi_{3}^{(1p,2p)}\rangle{} (57)

where

|Ψ3(1p,2p)\displaystyle|\Psi_{3}^{(1p,2p)}\rangle (Q1P2+Q2P1)|Ψ3=\displaystyle\equiv\left({{Q}_{{1}}{P}_{{2}}+{Q}_{{2}}{P}_{{1}}}\right)\ket{\Psi_{3}}=
=𝒩312(det~00de|de0+delt~000del|del\displaystyle=\frac{{\cal N}_{3}^{-1}}{2}\!\left(\sum_{de}\tilde{t}^{de}_{00}\ket{de0}+\sum_{del}\tilde{t}^{del}_{000}\ket{del}\right.
+dlmt~000dlm|dlm).\displaystyle\hskip 50.00008pt\left.+\sum_{dlm}\tilde{t}^{dlm}_{000}\ket{dlm}\right).{} (58)

Inspecting Eq. (VI.1) we note that the last term on the rhs is zero unless pd<2pFp_{d}<2p_{\text{F}}. It follows that this term must vanish if we to consider a very dilute AA-body system, i.e. the limit pF0p_{\text{F}}\to 0. Interestingly, in this limit also the first two terms coincide as t~000delt~000de0\tilde{t}^{del}_{000}\to\tilde{t}^{de0}_{000} with l0l\to 0. Hence, in the limit pF0p_{\text{F}}\to 0, Q3H^Q¯3|Ψ32𝒩31Q3H^T~^2|000{Q}_{{3}}\hat{H}\bar{Q}_{{3}}\ket{\Psi_{3}}\approx 2{\cal N}_{3}^{-1}{Q}_{{3}}\hat{H}\hat{\tilde{T}}_{{2}}\ket{000}. Recalling now that Eq. (45) is an asymptotic equation derived in the limit pa,pb,pcp_{a},p_{b},p_{c}\to\infty, and that in this limit T~^2T^2\hat{\tilde{T}}_{{2}}\to\hat{T}^{\infty}_{{2}}, we may conclude that for i,j,k=0i,j,k=0 the asymptotic 3-body cluster operator T^3\hat{T}^{\infty}_{{3}}, Eq. (VI), can be redefined replacing T^2\hat{T}^{\infty}_{{2}} with T~^2\hat{\tilde{T}}_{{2}}. The resulting operator admits

T^3|α3Q3|Ψ3\displaystyle\hat{T}^{\infty}_{{3}}\ket{\alpha_{3}}\approx{Q}_{{3}}\ket{\Psi_{3}}{} (59)

with |α32𝒩31|000\ket{\alpha_{3}}\equiv 2{\cal N}_{3}^{-1}\ket{000}.

Considering now the nuclear matter in the limit of dense matter (i.e. pFp_{\text{F}} very large compared to the Fermi momentum at saturation density), we have 12det~00de|de0T^2|000\frac{1}{2}\sum_{de}\tilde{t}^{de}_{00}\ket{de0}\to\hat{T}^{\infty}_{{2}}\ket{000}. Under this condition we also expect that the 2p1h2p1h terms 12delt~000del|del\frac{1}{2}\sum_{del}\tilde{t}^{del}_{000}\ket{del} are dominated by 2-body rather than 3-body correlations, i.e. most contributions will come from states with plpd,pep_{l}\ll p_{d},p_{e}. Therefore there is a 3-hole state |α3(2p)|\alpha_{3}^{(2p)}\rangle such that 12delt~000del|delT^2|α3(2p)\frac{1}{2}\sum_{del}\tilde{t}^{del}_{000}\ket{del}\approx\hat{T}^{\infty}_{{2}}|\alpha_{3}^{(2p)}\rangle. The 1p2h1p2h term 12dlmt~000dlm|dlm\frac{1}{2}\sum_{dlm}\tilde{t}^{dlm}_{000}\ket{dlm} is clearly zero if the momentum of state dd, pdp_{d}, is larger than 2pF2p_{\text{F}}. We also expect that the main contribution of this term will appear when pd,plpFp_{d},p_{l}\approx p_{\text{F}} and the 3rd momentum is approximately zero. Here again we can find a 0p3h0p3h state such that 12dlmt~000dlm|dlmT^2|α3(1p)\frac{1}{2}\sum_{dlm}\tilde{t}^{dlm}_{000}\ket{dlm}\approx\hat{T}^{\infty}_{{2}}|\alpha_{3}^{(1p)}\rangle. This observation implies that there is a 0p3h0p3h state α3\alpha_{3} such that

|Ψ3(1p,2p)T^2|α3,|\Psi_{3}^{(1p,2p)}\rangle\approx\hat{T}^{\infty}_{{2}}|\alpha_{3}\rangle, (60)

and hence also in this limit

T^3|α3Q3|Ψ3.\hat{T}^{\infty}_{{3}}\ket{\alpha_{3}}\approx{Q}_{{3}}\ket{\Psi_{3}}. (61)

This relation holds also for any value of pFp_{\text{F}} if we consider the most asymptotic high-momentum contribution to T^3\hat{T}^{\infty}_{{3}}, hence we expect it to approximately hold for finite nuclei as well.

Summarizing this discussion we conclude that, as in the 2-body case, (i) The asymptotic high-momentum behavior of T^3\hat{T}_{{3}} is related to a 3-body zero-energy eigen-function of the Schrödinger equation. (ii) This high momenta behavior is universal in the limited sense.

VII Summary

The CC method was utilized to set a more rigorous foundations for the successful GCF. To this end we have computed the 2- and 3-body cluster operators in the high momentum limit and showed that they act as the Bloch-Horowitz operators for the 2- and 3-body zero-energy eigen-states of the Schrödinger equation. We therefore concluded that the 2- and 3-body cluster operators in this high momentum limit are universal in the limited sense, that they do not depend on the system but do depend on the potential.

The presented method is systematic and opens up the path for including higher order corrections to the GCF. A more complete discussion regarding the asymptotic wave function factorization is postponed to a forthcoming article. We note that our results may be useful for general CC computations, as asymptotic expressions or approximations for the cluster operators.

Acknowledgements.
This research was supported by the ISRAEL SCIENCE FOUNDATION (grant No. 1086/21). The work of S. Beck was also supported by the Israel Ministry of Science and Technology (MOST). R. Weiss was supported by the Laboratory Directed Research and Development program of Los Alamos National Laboratory under project number 20210763PRD1.

Appendix A The asymptotics of T^2\hat{T}_{{2}}

To show that indeed the full 2-body amplitude T^2\hat{T}_{{2}} coincides with T^2\hat{T}^{\infty}_{{2}} in the limit pa,pbp_{a},p_{b}\to\infty we seek an iterative solution for Eq. (12)  [53]. To this end we denote by T^2(k)\hat{T}_{{2}}^{(k)} the approximate solution of T^2\hat{T}_{{2}} after kk iterations. Taking the asymptotic solution to be our initial guess T^2(0)=T^2\hat{T}_{{2}}^{(0)}=\hat{T}^{\infty}_{{2}}, we can write

T^2(k)=T^2+ΔT^2(1)+ΔT^2(2)++ΔT^2(k),\hat{T}_{{2}}^{(k)}=\hat{T}^{\infty}_{{2}}+\Delta\hat{T}^{{(1)}}_{2}+\Delta\hat{T}^{{(2)}}_{2}+\cdots+\Delta\hat{T}^{{(k)}}_{2}, (62)

where the kkth correction ΔT^2(k)\Delta\hat{T}^{{(k)}}_{2} is obtained by substituting T^2=T^2(k1)+ΔT^2(k)\hat{T}_{{2}}=\hat{T}_{{2}}^{(k-1)}+\Delta\hat{T}^{{(k)}}_{2} in (12) and solving the linearized equation.

The equation for ΔT^2(1)\Delta\hat{T}^{{(1)}}_{2} reads

0\displaystyle 0 =Φijab|V^+[H^0,T^2]+[H^0,ΔT^2(1)]+[V^,T^2]\displaystyle=\bra{\Phi_{ij}^{ab}}\hat{V}+[\hat{H}_{0},\hat{T}^{\infty}_{{2}}]+[\hat{H}_{0},\Delta\hat{T}^{{(1)}}_{2}]+[\hat{V},\hat{T}^{\infty}_{{2}}]
+[V^,ΔT^2(1)]+12[[V^,T^2],T^2]+[V^,T^3]+[V^,T^4]\displaystyle+[\hat{V},\Delta\hat{T}^{{(1)}}_{2}]+\frac{1}{2}[[\hat{V},\hat{T}^{\infty}_{{2}}],\hat{T}^{\infty}_{{2}}]+[\hat{V},\hat{T}_{{3}}]+[\hat{V},\hat{T}_{{4}}]
+12[[V^,T^2],ΔT^2(1)]+12[[V^,ΔT^2(1)],T^2]|Φ0\displaystyle+\frac{1}{2}[[\hat{V},\hat{T}^{\infty}_{{2}}],\Delta\hat{T}^{{(1)}}_{2}]+\frac{1}{2}[[\hat{V},\Delta\hat{T}^{{(1)}}_{2}],\hat{T}^{\infty}_{{2}}]\ket{\Phi_{0}}{} (63)

Utilizing eq. (32) we obtain

Δ\displaystyle\Delta tij(1)ab=(V^T^2)resEijab+EijEijab(t)ijab1EijabΦijab|\displaystyle t^{(1)ab}_{\quad ij}=-\frac{(\hat{V}\hat{T}^{\infty}_{{2}})_{\text{res}}}{E^{ab}_{ij}}+\frac{E^{ij}}{E^{ab}_{ij}}\left({t^{\infty}}\right)^{ab}_{ij}-\frac{1}{E^{ab}_{ij}}\bra{\Phi_{ij}^{ab}}
[V^,ΔT^2(1)]+12[[V^,T^2],T^2]+[V^,T^3]+[V^,T^4]\displaystyle\hskip 10.00002pt[\hat{V},\Delta\hat{T}^{{(1)}}_{2}]+\frac{1}{2}[[\hat{V},\hat{T}^{\infty}_{{2}}],\hat{T}^{\infty}_{{2}}]+[\hat{V},\hat{T}_{{3}}]+[\hat{V},\hat{T}_{{4}}]
+12[[V^,T^2],ΔT^2(1)]+12[[V^,ΔT^2(1)],T^2]|Φ0,\displaystyle\hskip 10.00002pt+\frac{1}{2}[[\hat{V},\hat{T}^{\infty}_{{2}}],\Delta\hat{T}^{{(1)}}_{2}]+\frac{1}{2}[[\hat{V},\Delta\hat{T}^{{(1)}}_{2}],\hat{T}^{\infty}_{{2}}]\ket{\Phi_{0}},{} (64)

where (V^T^2)res(\hat{V}\hat{T}^{\infty}_{{2}})_{\text{res}} stands for the terms that appear in (V) but are not included in (32). Asymptotically, as pa,pbp_{a},p_{b}\to\infty, the source terms should dominate

Δtij(1)ab\displaystyle\Delta t^{(1)ab}_{\quad ij}\to (V^T^2)resEijab+EijEijab(t)ijab1Eijab×\displaystyle-\frac{(\hat{V}\hat{T}^{\infty}_{{2}})_{\text{res}}}{E^{ab}_{ij}}+\frac{E^{ij}}{E^{ab}_{ij}}\left({t^{\infty}}\right)^{ab}_{ij}-\frac{1}{E^{ab}_{ij}}\times (65)
Φijab|12[[V^,T^2],T^2]+[V^,T^3]+[V^,T^4]|Φ0.\displaystyle\bra{\Phi_{ij}^{ab}}\frac{1}{2}[[\hat{V},\hat{T}^{\infty}_{{2}}],\hat{T}^{\infty}_{{2}}]+[\hat{V},\hat{T}_{{3}}]+[\hat{V},\hat{T}_{{4}}]\ket{\Phi_{0}}. (66)

Using the momentum arguments presented above and using the inherent hierarchy, the 3- and 4-body terms and (V^T^2)res(\hat{V}\hat{T}^{\infty}_{{2}})_{\text{res}} are suppressed by a factor V^/Eijab\langle\hat{V}\rangle/E^{ab}_{ij} (or Eij/EijabE^{ij}/E^{ab}_{ij}) compared to T^2\hat{T}^{\infty}_{{2}}. Hence asymptotically ΔT^2(1)T^2\Delta\hat{T}^{{(1)}}_{2}\ll\hat{T}^{\infty}_{{2}}. By iterating the process one can see that asymptotically the higher order k>1k>1 corrections are suppressed by a factor of the order (V^/Eijab)k(\langle\hat{V}\rangle/E^{ab}_{ij})^{k}. This completes the iterative proof that tijab(t)ijabt^{ab}_{ij}\to\left({t^{\infty}}\right)^{ab}_{ij}, and therefore in the high momentum limit we can replace the 2-body cluster operator T^2\hat{T}_{{2}} with the operator T^2\hat{T}^{\infty}_{{2}}.

Appendix B T^2\hat{T}^{\infty}_{{2}} as the Bloch-Horowitz operator

Recalling that G^0(E)=1E+iεH^0\hat{G}_{0}\left({E}\right)=\frac{1}{E+i\varepsilon-\hat{H}_{0}} and that it commutes with the projection operators Q2,P2{Q}_{{2}},{P}_{{2}} we can write for the Q2{Q}_{{2}} subspace

T^2\displaystyle\hat{T}^{\infty}_{{2}} =11Q2G^0(E)V^Q2G^0(E)H^P2\displaystyle=\frac{1}{1-{Q}_{{2}}\hat{G}_{0}\left({E}\right)\hat{V}}{Q}_{{2}}\hat{G}_{0}\left({E}\right)\hat{H}{P}_{{2}}
=1Q2G^0(E)Q2V^Q2G^0(E)Q2H^P2\displaystyle=\frac{1}{{Q}_{{2}}-\hat{G}_{0}\left({E}\right){Q}_{{2}}\hat{V}{Q}_{{2}}}\hat{G}_{0}\left({E}\right){Q}_{{2}}\hat{H}{P}_{{2}}
=1Q2(E+iεH^0V^)Q2Q2H^P2\displaystyle=\frac{1}{{Q}_{{2}}(E+i\varepsilon-\hat{H}_{0}-\hat{V}){Q}_{{2}}}{Q}_{{2}}\hat{H}{P}_{{2}}
=1Q2(E+iεH^)Q2Q2H^P2.\displaystyle=\frac{1}{{Q}_{{2}}(E+i\varepsilon-\hat{H}){Q}_{{2}}}{Q}_{{2}}\hat{H}{P}_{{2}}\;.{} (67)

Taking the value E=0E=0 it can be rewritten as

T^2=1Q2(0H^)Q2Q2H^P2.\displaystyle\hat{T}^{\infty}_{{2}}=\frac{1}{{Q}_{{2}}(0-\hat{H}){Q}_{{2}}}{Q}_{{2}}\hat{H}{P}_{{2}}\;.{} (68)

Appendix C The asymptotics of T^3\hat{T}_{{3}}

To show that tijkabc(t)ijkabct^{abc}_{ijk}\to\left({t^{\infty}}\right)_{ijk}^{abc} as pa,pb,pcp_{a},p_{b},p_{c}\to\infty we substitute T^3=T^3+ΔT^3\hat{T}_{{3}}=\hat{T}^{\infty}_{{3}}+\Delta\hat{T}_{3} in eq. (14) and solve for ΔT^3\Delta\hat{T}_{3} after using the definition of T^3\hat{T}^{\infty}_{{3}} in eq. (VI). Moreover, as explained at section IV, 12[[V^,T^2],T^2][V^,T^2]\frac{1}{2}[[\hat{V},\hat{T}_{{2}}],\hat{T}_{{2}}]\ll[\hat{V},\hat{T}_{{2}}] in the limit pa,pb,pcp_{a},p_{b},p_{c}\to\infty, hence the equation for Δtijkabc\Delta t^{abc}_{ijk} becomes

Δ\displaystyle\Delta tijkabc=(V^T^3)resEijkabc+EijkEijkabc(t)ijkabc\displaystyle t^{abc}_{ijk}=-\frac{(\hat{V}\hat{T}^{\infty}_{{3}})_{\text{res}}}{E^{abc}_{ijk}}+\frac{E^{ijk}}{E^{abc}_{ijk}}\left({t^{\infty}}\right)^{abc}_{ijk}
1EijkabcΦijkabc|[H^0,ΔT^3]+[V^,ΔT^2]+[V^,ΔT^3]\displaystyle-\frac{1}{E^{abc}_{ijk}}\bra{\Phi_{ijk}^{abc}}[\hat{H}_{0},\Delta\hat{T}_{3}]+[\hat{V},\Delta\hat{T}_{2}]+[\hat{V},\Delta\hat{T}_{3}]
+12[[V^,T^2],ΔT^3]+[V^,T^4]+[V^,T^5]|Φ0\displaystyle+\frac{1}{2}[[\hat{V},\hat{T}_{{2}}],\Delta\hat{T}_{3}]+[\hat{V},\hat{T}_{{4}}]+[\hat{V},\hat{T}_{{5}}]\ket{\Phi_{0}}\;{} (69)

where (V^T^3)res(\hat{V}\hat{T}^{\infty}_{{3}})_{\text{res}} stands for the terms that appear in (VI) but are not included in (45) and ΔT^2=T^2T^2\Delta\hat{T}_{2}=\hat{T}_{{2}}-\hat{T}^{\infty}_{{2}}. Asymptotically, the source terms should dominate, and thus

Δtijkabc\displaystyle\Delta t^{abc}_{ijk}\to (V^T^3)resEijkabc+EijkEijkabc(t)ijkabc1Eijkabc×\displaystyle-\frac{(\hat{V}\hat{T}^{\infty}_{{3}})_{\text{res}}}{E^{abc}_{ijk}}+\frac{E^{ijk}}{E^{abc}_{ijk}}\left({t^{\infty}}\right)^{abc}_{ijk}-\frac{1}{E^{abc}_{ijk}}\times
Φijkabc|[V^,ΔT^2]+[V^,T^4]+[V^,T^5]|Φ0.\displaystyle\bra{\Phi_{ijk}^{abc}}[\hat{V},\Delta\hat{T}_{2}]+[\hat{V},\hat{T}_{{4}}]+[\hat{V},\hat{T}_{{5}}]\ket{\Phi_{0}}\;.{} (70)

The terms in the first row are trivially much smaller than (t)ijkabc(t^{\infty})^{abc}_{ijk}. Also, the 4- and 5-body terms are also much smaller than (t)ijkabc(t^{\infty})^{abc}_{ijk} due to hierarchy and the suppression of the factor 1Eijkabc\frac{1}{E^{abc}_{ijk}}. For the term [V^,ΔT^2][\hat{V},\Delta\hat{T}_{2}] we can use the results of the previous section (Δt)jkcd(t)jkcd\left({\Delta t}\right)^{cd}_{jk}\ll\left({t^{\infty}}\right)^{cd}_{jk} and note, from eq. (VI), that (t)ijkabc1Eijkabc𝒮^abc[𝒮^ijk[Vidab(t)jkcd]]\left({t^{\infty}}\right)^{abc}_{ijk}\sim-\frac{1}{E^{abc}_{ijk}}{\cal\hat{S}}_{abc}\left[{{\cal\hat{S}}_{ijk}\left[{V^{ab}_{id}\left({t^{\infty}}\right)^{cd}_{jk}}\right]}\right]. Altogether we get the desired reuslt (Δt)ijkabc(t)ijkabc\left({\Delta t}\right)^{abc}_{ijk}\ll\left({t^{\infty}}\right)^{abc}_{ijk}, i.e. tijkabc(t)ijkabct^{abc}_{ijk}\to\left({t^{\infty}}\right)^{abc}_{ijk}.

References

  • Ciofi degli Atti [2015] C. Ciofi degli Atti, In-medium short-range dynamics of nucleons: Recent theoretical and experimental advances, Phys. Rept. 590, 1 (2015).
  • Hen et al. [2017] O. Hen, G. A. Miller, E. Piasetzky, and L. B. Weinstein, Nucleon-Nucleon Correlations, Short-lived Excitations, and the Quarks Within, Rev. Mod. Phys. 89, 045002 (2017).
  • Frankfurt and Strikman [1981] L. L. Frankfurt and M. I. Strikman, High-energy phenomena, short-range nuclear structure and qcd, Phys. Rep. 76, 215 (1981).
  • Frankfurt and Strikman [1988] L. Frankfurt and M. Strikman, Hard nuclear processes and microscopic nuclear structure, Phys. Rep. 160, 235 (1988).
  • Tang et al. [2003] A. Tang et al., n-p short range correlations from (p,2p + n) measurements, Phys. Rev. Lett. 90, 042301 (2003)arXiv:nucl-ex/0206003 .
  • Piasetzky et al. [2006] E. Piasetzky, M. Sargsian, L. Frankfurt, M. Strikman, and J. W. Watson, Evidence for strong dominance of proton-neutron correlations in nuclei, Phys. Rev. Lett. 97, 162504 (2006).
  • Shneor et al. [2007] R. Shneor et al. (Jefferson Lab Hall A), Investigation of proton-proton short-range correlations via the C-12(e, e-prime pp) reaction, Phys. Rev. Lett. 99, 072501 (2007)arXiv:nucl-ex/0703023 .
  • Subedi et al. [2008] R. Subedi et al., Probing Cold Dense Nuclear Matter, Science 320, 1476 (2008)arXiv:0908.1514 [nucl-ex] .
  • Korover et al. [2014] I. Korover, N. Muangma, O. Hen, et al., Probing the Repulsive Core of the Nucleon-Nucleon Interaction via the 4He(e,e’pN) Triple-Coincidence Reaction, Phys. Rev. Lett. 113, 022501 (2014).
  • Cohen et al. [2018] E. O. Cohen et al. (CLAS Collaboration), Center of Mass Motion of Short-Range Correlated Nucleon Pairs studied via the A(e,epp)A(e,e^{\prime}pp) Reaction, Phys. Rev. Lett. 121, 092501 (2018)arXiv:1805.01981 [nucl-ex] .
  • Hen et al. [2014] O. Hen et al., Momentum sharing in imbalanced Fermi systems, Science 346, 614 (2014)arXiv:1412.0138 [nucl-ex] .
  • Duer et al. [2019] M. Duer et al. (CLAS Collaboration), Direct Observation of Proton-Neutron Short-Range Correlation Dominance in Heavy Nuclei, Phys. Rev. Lett. 122, 172502 (2019)arXiv:1810.05343 [nucl-ex] .
  • Schmidt et al. [2020] A. Schmidt et al. (CLAS), Probing the core of the strong nuclear interaction, Nature 578, 540 (2020)arXiv:2004.11221 [nucl-ex] .
  • Korover et al. [2021] I. Korover, J. Pybus, A. Schmidt, F. Hauenstein, M. Duer, O. Hen, E. Piasetzky, L. Weinstein, D. Higinbotham, S. Adhikari, K. Adhikari, M. Amaryan, G. Angelini, H. Atac, L. Barion, M. Battaglieri, A. Beck, I. Bedlinskiy, F. Benmokhtar, A. Bianconi, A. Biselli, S. Boiarinov, W. Briscoe, W. Brooks, D. Bulumulla, V. Burkert, D. Carman, A. Celentano, P. Chatagnon, T. Chetry, L. Clark, B. Clary, P. Cole, M. Contalbrigo, V. Crede, R. Cruz-Torres, A. D’Angelo, R. De Vita, M. Defurne, A. Denniston, A. Deur, S. Diehl, C. Djalali, R. Dupre, H. Egiyan, M. Ehrhart, A. El Alaoui, L. El Fassi, L. Elouadrhiri, P. Eugenio, R. Fersch, A. Filippi, T. Forest, G. Gavalian, F. Girod, E. Golovatch, R. Gothe, K. Griffioen, M. Guidal, K. Hafidi, H. Hakobyan, N. Harrison, M. Hattawy, T. Hayward, D. Heddle, K. Hicks, M. Holtrop, Y. Ilieva, D. Ireland, E. Isupov, D. Jenkins, H. Jo, K. Joo, S. Joosten, D. Keller, M. Khachatryan, A. Khanal, M. Khandaker, A. Kim, C. Kim, F. Klein, V. Kubarovsky, L. Lanza, M. Leali, P. Lenisa, K. Livingston, V. Lucherini, I. MacGregor, D. Marchand, N. Markov, L. Marsicano, V. Mascagna, B. McKinnon, S. Mey-Tal Beck, T. Mineeva, M. Mirazita, A. Movsisyan, C. Munoz Camacho, B. Mustapha, P. Nadel-Turonski, K. Neupane, G. Niculescu, M. Osipenko, A. Ostrovidov, M. Paolone, L. Pappalardo, R. Paremuzyan, E. Pasyuk, W. Phelps, O. Pogorelko, J. Price, Y. Prok, D. Protopopescu, B. Raue, M. Ripani, J. Ritman, A. Rizzo, G. Rosner, J. Rowley, F. Sabatié, C. Salgado, R. Schumacher, E. Segarra, Y. Sharabian, U. Shrestha, D. Sokhan, O. Soto, N. Sparveris, S. Stepanyan, I. Strakovsky, S. Strauch, J. Tan, N. Tyler, M. Ungaro, L. Venturelli, H. Voskanyan, E. Voutier, T. Wang, D. Watts, X. Wei, M. Wood, N. Zachariou, J. Zhang, Z. Zhao, and X. Zheng, 12c(e,e’pn) measurements of short range correlations in the tensor-to-scalar interaction transition region, Physics Letters B 820, 136523 (2021).
  • Frankfurt et al. [1993] L. Frankfurt, M. Strikman, D. Day, and M. Sargsyan, Evidence for short-range correlations from high q2 (e,e’) reactions, Phys. Rev. C 48, 2451 (1993).
  • Egiyan et al. [2003] K. Egiyan et al. (CLAS Collaboration), Phys. Rev. C 68, 014313 (2003).
  • Egiyan et al. [2006] K. Egiyan et al. (CLAS Collaboration), Measurement of 2- and 3-nucleon short range correlation probabilities in nuclei, Phys. Rev. Lett. 96, 082501 (2006).
  • Fomin et al. [2012] N. Fomin et al., New measurements of high-momentum nucleons and short-range structures in nuclei, Phys. Rev. Lett. 108, 092502 (2012).
  • Schmookler et al. [2019] B. Schmookler et al. (CLAS Collaboration), Modified structure of protons and neutrons in correlated pairs, Nature 566, 354 (2019).
  • Feldmeier et al. [2011] H. Feldmeier, W. Horiuchi, T. Neff, and Y. Suzuki, Universality of short-range nucleon-nucleon correlations, Phys. Rev. C 84, 054003 (2011)arXiv:1107.4956 [nucl-th] .
  • Alvioli et al. [2013a] M. Alvioli, C. Ciofi degli Atti, L. P. Kaptari, C. B. Mezzetti, and H. Morita, Nucleon momentum distributions, their spin-isospin dependence and short-range correlations, Phys. Rev. C87, 034603 (2013a)arXiv:1211.0134 [nucl-th] .
  • Wiringa et al. [2014] R. B. Wiringa, R. Schiavilla, S. C. Pieper, and J. Carlson, Nucleon and nucleon-pair momentum distributions in a12a\leq 12, Phys. Rev. C 89, 024305 (2014).
  • Rios et al. [2014] A. Rios, A. Polls, and W. H. Dickhoff, Density and isospin asymmetry dependence of high-momentum components, Phys. Rev. C89, 044303 (2014)arXiv:1312.7307 [nucl-th] .
  • Ryckebusch et al. [2019] J. Ryckebusch, W. Cosyn, T. Vieijra, and C. Casert, Isospin composition of the high-momentum fluctuations in nuclei from asymptotic momentum distributions, Phys. Rev. C 100, 054620 (2019)arXiv:1907.07259 [nucl-th] .
  • Alvioli et al. [2013b] M. Alvioli, C. Ciofi Degli Atti, L. P. Kaptari, C. B. Mezzetti, and H. Morita, Universality of nucleon-nucleon short-range correlations and nucleon momentum distributions, Int. J. Mod. Phys. E22, 1330021 (2013b)arXiv:1306.6235 [nucl-th] .
  • Neff et al. [2015] T. Neff, H. Feldmeier, and W. Horiuchi, Short-range correlations in nuclei with similarity renormalization group transformations, Phys. Rev. C 92, 024003 (2015).
  • Ryckebusch et al. [2015] J. Ryckebusch, M. Vanhalst, and W. Cosyn, Stylized features of single-nucleon momentum distributions, Journal of Physics G: Nuclear and Particle Physics 42, 055104 (2015).
  • Alvioli et al. [2008] M. Alvioli, C. Ciofi degli Atti, and H. Morita, Proton-neutron and proton-proton correlations in medium-weight nuclei and the role of the tensor force, Phys. Rev. Lett. 100, 162503 (2008).
  • Ye et al. [2018] Z. Ye et al. (The Jefferson Lab Hall A Collaboration), Search for three-nucleon short-range correlations in light nuclei, Phys. Rev. C 97, 065204 (2018).
  • Sargsian et al. [2019] M. M. Sargsian, D. B. Day, L. L. Frankfurt, and M. I. Strikman, Searching for three-nucleon short-range correlations, Phys. Rev. C 100, 044320 (2019).
  • Tan [2008a] S. Tan, Energetics of a strongly correlated fermi gas, Annals of Physics 323, 2952 (2008a).
  • Tan [2008b] S. Tan, Large momentum part of a strongly correlated fermi gas, Annals of Physics 323, 2971 (2008b).
  • Tan [2008c] S. Tan, Generalized virial theorem and pressure relation for a strongly correlated fermi gas, Annals of Physics 323, 2987 (2008c).
  • Braaten [2012] E. Braaten, Universal relations for fermions with large scattering length, in The BCS-BEC Crossover and the Unitary Fermi Gas, edited by W. Zwerger (Springer, Berlin, 2012).
  • Weiss et al. [2015a] R. Weiss, B. Bazak, and N. Barnea, Nuclear neutron-proton contact and the photoabsorption cross section, Phys. Rev. Lett. 114, 012501 (2015a).
  • Weiss et al. [2015b] R. Weiss, B. Bazak, and N. Barnea, Generalized nuclear contacts and momentum distributions, Phys. Rev. C92, 054311 (2015b)arXiv:1503.07047 [nucl-th] .
  • Weiss et al. [2016a] R. Weiss, E. Pazy, and N. Barnea, Short range correlations: The important role of few-body dynamics in many-body systems, Few-Body Systems 5810.1007/s00601-016-1165-2 (2016a).
  • Weiss and Barnea [2017] R. Weiss and N. Barnea, Contact formalism for coupled channels, Phys. Rev. C 96, 041303 (2017).
  • Weiss et al. [2018] R. Weiss, R. Cruz-Torres, N. Barnea, E. Piasetzky, and O. Hen, The nuclear contacts and short range correlations in nuclei, Phys. Lett. B 780, 211 (2018).
  • Cruz-Torres et al. [2020] R. Cruz-Torres, D. Lonardoni, R. Weiss, M. Piarulli, N. Barnea, D. W. Higinbotham, E. Piasetzky, A. Schmidt, L. B. Weinstein, R. B. Wiringa, and O. Hen, Many-body factorization and position-momentum equivalence of nuclear short-range correlations, Nature Physics  (2020)arXiv:1907.03658 [nucl-th] .
  • Weiss et al. [2016b] R. Weiss, B. Bazak, and N. Barnea, The generalized nuclear contact and its application to the photoabsorption cross-section, The European Physical Journal A 5210.1140/epja/i2016-16092-3 (2016b).
  • Weiss et al. [2019a] R. Weiss, I. Korover, E. Piasetzky, O. Hen, and N. Barnea, Energy and momentum dependence of nuclear short-range correlations - Spectral function, exclusive scattering experiments and the contact formalism, Phys. Lett. B791, 242 (2019a)arXiv:1806.10217 [nucl-th] .
  • Weiss et al. [2019b] R. Weiss, A. Schmidt, G. A. Miller, and N. Barnea, Short-range correlations and the charge density, Physics Letters B 790, 484 (2019b).
  • Pybus et al. [2020] J. Pybus, I. Korover, R. Weiss, A. Schmidt, N. Barnea, D. Higinbotham, E. Piasetzky, M. Strikman, L. Weinstein, and O. Hen, Generalized contact formalism analysis of the 4He(e,epN)(e,e^{\prime}pN) reaction, Phys. Lett. B 805, 135429 (2020)arXiv:2003.02318 [nucl-th] .
  • Patsyuk et al. [2021] M. Patsyuk, J. Kahlbow, G. Laskaris, M. Duer, V. Lenivenko, E. P. Segarra, T. Atovullaev, G. Johansson, T. Aumann, A. Corsi, and et al., Unperturbed inverse kinematics nucleon knockout measurements with a carbon beam, Nature Physics 17, 693–699 (2021).
  • Weiss et al. [2021a] R. Weiss, A. W. Denniston, J. R. Pybus, O. Hen, E. Piasetzky, A. Schmidt, L. B. Weinstein, and N. Barnea, Extracting the number of short-range correlated nucleon pairs from inclusive electron scattering data, Phys. Rev. C 103, L031301 (2021a).
  • Weiss et al. [2021b] R. Weiss, P. Soriano, A. Lovato, J. Menendez, and R. B. Wiringa, Neutrinoless double-beta decay: combining quantum monte carlo and the nuclear shell model with the generalized contact formalism (2021b), arXiv:2112.08146 [nucl-th] .
  • Anderson et al. [2010] E. R. Anderson, S. K. Bogner, R. J. Furnstahl, and R. J. Perry, Operator evolution via the similarity renormalization group: The deuteron, Phys. Rev. C 82, 054001 (2010).
  • Bogner and Roscher [2012] S. K. Bogner and D. Roscher, High-momentum tails from low-momentum effective theories, Phys. Rev. C 86, 064304 (2012).
  • Tropiano et al. [2021] A. J. Tropiano, S. K. Bogner, and R. J. Furnstahl, Short-range correlation physics at low renormalization group resolution, Phys. Rev. C 104, 034311 (2021).
  • Bartlett and Musiał [2007] R. J. Bartlett and M. Musiał, Coupled-cluster theory in quantum chemistry, Rev. Mod. Phys. 79, 291 (2007).
  • Hagen et al. [2014a] G. Hagen, T. Papenbrock, M. Hjorth-Jensen, and D. J. Dean, Coupled-cluster computations of atomic nuclei, Reports on Progress in Physics 77, 096302 (2014a).
  • I. Shavitt [2009] R. J. B. I. Shavitt, Many-Body Methods in Chemistry and Physics MBPT and Coupled-Cluster Theory (Cambridge Molecular Science, 2009).
  • Baardsen et al. [2013] G. Baardsen, A. Ekström, G. Hagen, and M. Hjorth-Jensen, Coupled-cluster studies of infinite nuclear matter, Phys. Rev. C 88, 054312 (2013).
  • Hagen et al. [2014b] G. Hagen, T. Papenbrock, A. Ekström, K. A. Wendt, G. Baardsen, S. Gandolfi, M. Hjorth-Jensen, and C. J. Horowitz, Coupled-cluster calculations of nucleonic matter, Phys. Rev. C 89, 014319 (2014b).
  • Bishop and Lührmann [1978] R. F. Bishop and K. H. Lührmann, Electron correlations: I. ground-state results in the high-density regime, Phys. Rev. B 17, 3757 (1978).
  • Amado and Woloshyn [1976] R. Amado and R. Woloshyn, Momentum distributions in the nucleus, Physics Letters B 62, 253 (1976).
  • Zabolitzky and Ey [1978] J. Zabolitzky and W. Ey, Momentum distributions of nucleons in nuclei, Physics Letters B 76, 527 (1978).
  • Bloch and Horowitz [1958] C. Bloch and J. Horowitz, Sur la détermination des premiers états d’un système de fermions dans le cas dégénéré, Nuclear Physics 8, 91 (1958).