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

Particle Interferometry in a Moat Regime

Fabian Rennecke [email protected] Institute for Theoretical Physics, Justus Liebig University Giessen, Heinrich-Buff-Ring 16, 35392 Giessen, Germany Helmholtz Research Academy Hesse for FAIR (HFHF), Campus Giessen, Giessen, Germany    Robert D. Pisarski [email protected] Department of Physics, Brookhaven National Laboratory, Upton, New York 11973, USA    Dirk H. Rischke [email protected] Institute for Theoretical Physics, Goethe University, Max-von-Laue-Str. 1, 60438 Frankfurt am Main, Germany Helmholtz Research Academy Hesse for FAIR (HFHF), Campus Riedberg, Frankfurt am Main, Germany
Abstract

Dense strongly interacting matter can exhibit regimes with spatial modulations, akin to crystalline phases. In this case particles can have a moat spectrum with minimal energy at nonzero momentum. We show that particle interferometry is a sensitive probe of such a regime in heavy-ion collisions. To this end, we develop a field-theoretical formalism that relates particle spectra to in-medium real-time correlation functions of quantum fields on curved hypersurfaces of spacetime. This is then applied to the study of Bose-Einstein correlations in a moat regime in heavy-ion collisions. The resulting two-particle spectra exhibit peaks at nonzero average pair momentum, in contrast to the two-particle spectra in a normal phase, which peak at zero momentum. These peaks lead to non-trivial structures in the ratio of two-particle correlation functions, which should be experimentally measurable if the resolution in the direction of average pair momentum is sufficiently large. We propose these structures in the correlation-function ratios as clear signature of a moat regime and spatially modulated phases in quantum chromodynamics (QCD).

I Introduction

Understanding the phase structure of QCD is a long-standing open problem of fundamental physics. Lattice QCD [1, 2, 3] and functional continuum methods [4, 5, 6] have by now established that QCD at vanishing and small densities has a crossover transition from hadrons at low temperature to deconfined quarks and gluons at very high temperature. At intermediate temperatures, matter could be in a ‘stringy fluid’ state [7]. For extremely high densities, well above any density found in Nature, it is known that at low temperatures QCD is in a color-superconducting state [8, 9, 10]. At intermediate densities there are presently no first-principles results available, but a wealth of different phases have been predicted by model studies [11].

Strongly interacting matter in this intermediate density region is of particular interest, as it is produced in heavy-ion collisions and is present in neutron stars [12]. Among the phases suggested by model studies are such with a spatial modulation of a certain quantity, like the density or the order parameter for chiral symmetry breaking or color superconductivity. These could be crystalline or inhomogeneous phases [11, 13], liquid-crystal like phases [14, 15, 16], or a quantum pion liquid [17, 18]. Where and if at all these phases are realized in the phase diagram is ultimately decided by long-range quantum fluctuations, and understanding them therefore poses a great challenge. However, a characteristic, or at least precursory, feature of all these phases is the occurrence of a moat energy spectrum [17, 19, 20]. In a low-momentum expansion, the energy of a relativistic particle (assuming rotational invariance) is

E𝕡=z𝕡2+w𝕡4+m2.\displaystyle E_{\mathbb{p}}=\sqrt{z\,\mathbb{p}^{2}+w\,\mathbb{p}^{4}+m^{2}}\,. (1)

A free relativistic particle has z=1z=1 and w=0w=0 and thus a minimal energy at rest, 𝕡min=0\mathbb{p}_{\rm min}=0. In contrast, in the presence of spatial modulations with a wavenumber 𝕜00\mathbb{k}_{0}\neq 0, the energy of particles is minimized at nonzero momentum, 𝕡min=𝕜0\mathbb{p}_{\rm min}=\mathbb{k}_{0}. At low momenta, their energy is described by Eq. (1) with z<1z<1 and, for stability, w>0w>0. The energy then looks like a moat. We refer to the region of the phase diagram where particles have such an energy as a moat regime.

Since a moat regime occurs whenever spatial modulations are present, but is to some extent agnostic to which type of phase is actually realized, it is both theoretically and, as we will show here, also experimentally more easily accessible. The moat regime can be viewed as an indicator for the presence of spatially modulated phases in QCD. First theoretical evidence that this is indeed the case has been found in Ref. [5]. Stimulated by the first suggestions in Refs. [21, 19], we will show how a moat regime can be detected in heavy-ion collisions.

Refer to caption
Figure 1: Instead of describing correlations at equal time on a flat hypersurface of spacetime, where the involved particles can be subject to different thermodynamic conditions (left figure), we describe them on a curved hypersurface defined by fixed thermodynamic conditions (right figure).
Refer to caption
Refer to caption
Refer to caption
Figure 2: Ratios of normalized two-particle correlations, C(P,ΔPout)C(P,ΔPlong)\frac{C(P,\Delta P_{\rm out})}{C(P,\Delta P_{\rm long})}, C(P,ΔPout)C(P,ΔPside)\frac{C(P,\Delta P_{\rm out})}{C(P,\Delta P_{\rm side})} and C(P,ΔPside)C(P,ΔPlong)\frac{C(P,\Delta P_{\rm side})}{C(P,\Delta P_{\rm long})} in the normal phase. Out, side, and long denote the direction of the relative momentum of the particle pair, ΔP=|Δ|\Delta P=|\mathbb{\Delta P}|, relative to the average pair momentum P=||P=|\mathbb{P}|. In the standard notation, the average pair momentum PKP\to K, and the relative momentum ΔPq\Delta P\to q.
Refer to caption
Refer to caption
Refer to caption
Figure 3: The same ratios as in Fig. 2, but in a moat regime. The additional peaks around P100P\approx 100 MeV reflect the wave number of the underlying spatial modulation and are therefore clear signatures of a moat regime.

The basic idea is that since particles in a moat regime have minimal energy at nonzero momentum 𝕡min\mathbb{p}_{\rm min}, particle production is enhanced at this momentum. It has been shown in Ref. [19] that this leads to a peak in the single-particle spectrum which is directly related to 𝕡min\mathbb{p}_{\rm min}, and that such a peak persists in correlations that are generated by thermodynamic fluctuations. Another prominent source of correlations, which so far has not been studied in this context, is due to the quantum statistics of identical particles. The study of these correlations is known as Hanbury-Brown–Twiss (HBT) interferometry, and was originally used to infer the angular diameter of astronomical objects [22]. In particle physics, HBT correlations through interference measurements of pions were first studied in Ref. [23]. Since then, identical-particle interferometry has been developed into the standard tool for the extraction of information on the spacetime structure of the medium created in heavy-ion collisions [24, 25, 26, 27, 28].

Since pions are expected to be sensitive to a moat regime [5, 21], we propose pion interferometry as a way to detect such a regime in heavy-ion collisions. Based on exploratory results in QCD [5], we expect such a regime in the intermediate density region with μBT4\frac{\mu_{B}}{T}\gtrsim 4, where μB\mu_{B} is the chemical potential of baryons and TT the temperature. This is within reach of present and future heavy-ion experiments at lower beam energies, such as STAR at RHIC, HADES at GSI, CBM at FAIR, NA61/SHINE at CERN [29, 30, 31].

This paper is organized as follows: We summarize our main results in Sec. II. In Sec. III we develop the formalism to describe particle correlations in a moat regime. This requires techniques of quantum field theory in curved spacetime, which are discussed in Sec. III.1. We use this to relate 2n2n-point functions of quantum fields to nn-particle spectra in Sec. III.2, with a special focus on the two-particle spectra relevant for the present purposes. This formalism is then applied to the study of HBT correlations in Sec. IV. We first demonstrate in Sec. IV.1 that the relevant effects are accessible through four-point functions in the Gaussian approximation, and then specify the two-point function of a quasiparticle in a moat regime, which is required as an input. Then, in Sec. IV.3, we set up a simple model for the energy spectrum and the hypersurface on which the correlations are studied, and numerically analyze the resulting single- and two-particle spectra, as well as the HBT radii in Sec. IV.4. We conclude in Sec. V. In a short appendix, App. A, we derive the contribution of particle-antiparticle interference to the two-particle spectrum.

II Summary of our results

In order to describe particle spectra in heavy-ion collisions, we first develop a formalism that allows us to compute correlations of particles on curved hypersurfaces in spacetime. This is necessary since, owing to the collision dynamics, fixed thermodynamic conditions, such as a given temperature and density, are only present on hypersurfaces which in general differ from spatial 3\mathbb{R}^{3}. This is illustrated in Fig. 1. While this idea underlies the well-known Cooper-Frye formalism for single-particle spectra [32], and the hydrodynamic description of two-particle correlations [33], only spectra of free particles can be described by existing frameworks. In order to study signatures of the QCD phase structure, we need to describe correlations of particles which are subject to in-medium modifications. Based on heuristic arguments, a first attempt to do this for single-particle spectra has been made in Ref. [19], but a microscopic derivation has been lacking. This is provided here. The formalism we develop is based on the quantization of quasiparticles in curved spacetime and their real-time correlation functions. Using a Gaussian approximation, we explicitly derive a relation between the two-particle spectrum and the statistical and spectral function of the particles. This allows us to systematically incorporate in-medium modifications.

Using this formalism, we show that, in local thermodynamic equilibrium, the in-medium properties of particles contribute to the correlations generated by interference through the spectral functions of the particles evaluated at the average momentum of the particle pair. Hence, the nontrivial information about the phase structure is encoded in the average-momentum dependence of two-particle correlations. In experimental studies of particle interference it is customary to normalize these correlations,

C(,Δ)=n2(,Δ)n1(+12Δ)n1(12Δ),\displaystyle C(\mathbb{P},\mathbb{\Delta P})=\frac{n_{2}(\mathbb{P},\mathbb{\Delta P})}{n_{1}\big{(}\mathbb{P}+\frac{1}{2}\mathbb{\Delta P}\big{)}\,n_{1}\big{(}\mathbb{P}-\frac{1}{2}\mathbb{\Delta P}\big{)}}\,, (2)

where n1n_{1} and n2n_{2} are the single- and two-particle spectra and \mathbb{P} and Δ\mathbb{\Delta P} the average and relative momenta of the particle pair. Our main results are shown in Figs. 2 and 3. They show ratios of normalized two-particle correlations in three different directions for the relative pair momentum relative to the average momentum, called out, side, and long [34, 35]. Figure 2 shows the ratios for particles in the normal phase, i.e. where their energy follows Eq. (1) with z=1z=1 and w=0w=0. Figure 2 shows the same ratios, but in a moat regime. In the moat regime, the ratios show pronounced peaks which are absent in the normal phase. The location of these peaks is determined by the momentum which minimizes the energy of the particles. This momentum, in turn, is related to the wave number of the underlying spatial modulation. If the experimental resolution in the average pair momentum is sufficiently large, these peaks should be experimentally detectable and thus provide a clear signal for a moat regime and, thus, spatially modulated phases in the QCD phase diagram. We note that we use illustrative models to describe the moat regime in this work, so the position of the peak can vary quantitatively depending on the value of 𝕡min0\mathbb{p}_{\rm min}\neq 0, but its existence is nevertheless a robust qualitative prediction of our study. This is the main result of the present work.

III Particle spectra on a hypersurface

We pointed out in Ref. [19] that particle spectra in a moat regime cannot be addressed by using the conventional Cooper-Frye formalism [32]. This applies to all sorts of particle-number correlations, not only the ones generated by thermodynamic fluctuations studied in Ref. [19]. To be most flexible, we therefore aim for a field-theoretic expression of particle spectra in terms of elementary correlations of quantum fields on a (freeze-out) hypersurface Σ\Sigma. This is reminiscent of the LSZ reduction formalism.

Particle spectra are related to the phase-space density of the particle number. As such, unless particle number is conserved for the particles under consideration, this notion is only well defined in an ‘asymptotic’ sense. This specifically means that if it is possible to define ladder operators a𝕡a_{\mathbb{p}} and a𝕡a_{\mathbb{p}}^{\dagger}, which annihilate and create a single-particle state with spatial momentum 𝕡\mathbb{p} and energy ω𝕡\omega_{\mathbb{p}}, then, e.g., the covariant single- and two-particle spectra are given by

n1(𝕡)=(2π)3ω𝕡dN1d3p=ω𝕡a𝕡a𝕡,n2(𝕡,𝕢)=(2π)6ω𝕡ω𝕢dN2d3pd3q=ω𝕡ω𝕢a𝕡a𝕢a𝕡a𝕢,\displaystyle\begin{split}n_{1}(\mathbb{p})=(2\pi)^{3}\omega_{\mathbb{p}}\frac{dN_{1}}{d^{3}p}&=\omega_{\mathbb{p}}\big{\langle}a_{\mathbb{p}}^{\dagger}a_{\mathbb{p}}\big{\rangle}\,,\\ n_{2}(\mathbb{p},\mathbb{q})=(2\pi)^{6}\omega_{\mathbb{p}}\omega_{\mathbb{q}}\frac{dN_{2}}{d^{3}p\,d^{3}q}&=\omega_{\mathbb{p}}\omega_{\mathbb{q}}\big{\langle}a_{\mathbb{p}}^{\dagger}a_{\mathbb{q}}^{\dagger}a_{\mathbb{p}}a_{\mathbb{q}}\big{\rangle}\,,\end{split} (3)

and analogously for further multi-particle spectra. The nn-particle spectra for n2n\geq 2 obviously also encode particle-number correlations. Note that the expectation values do not necessarily have to be vacuum expectation values.

The averages should be with respect to the density matrix on the hypersurface. This can be most conveniently done by quantizing the system on the hypersurface. Another possibility would be to quantize in the rest frame and then boost onto the surface. However, if the system lacks boost invariance, this can be problematic.

One way to compute these spectra is using incoherent sources as discussed, e.g., in Refs. [24, 27]. Here, we aim at computing these correlations on a hypersurface of the fireball created in heavy-ion collisions and therefore seek to find a formalism more suited to do this directly. To this end, we will develop a new formalism to study spectra on general hypersurfaces.

III.1 Quasiparticles in foliated spacetime

As illustrated in Fig. 1, in situations like heavy-ion collisions, a fixed temperature, for example, is typically found on a three-dimensional hypersurface Σ\Sigma of 4dd Minkowski spacetime which differs from flat 3\mathbb{R}^{3}. Thus, instead of describing these particles in Minkowski space, it is advantageous to describe them in terms of an appropriate foliation of spacetime. To this end, we use that any hypersurface Σ\Sigma can be defined by parametric equations of the form

xμ=xμ(wi),\displaystyle x^{\mu}=x^{\mu}(w^{i})\,, (4)

where wi(x)w^{i}(x) with i=1,2,3i=1,2,3 are coordinates intrinsic to Σ\Sigma [36]. We define tangent vectors pointing into ‘spatial’ directions,

eiμ=xμwi,\displaystyle e_{i}^{\mu}=\frac{\partial x^{\mu}}{\partial w^{i}}\,, (5)

which give rise to the induced metric (first fundamental form) on Σ\Sigma,

Gij=gμνeiμejν,\displaystyle G_{ij}=-g_{\mu\nu}\,e_{i}^{\mu}e_{j}^{\nu}\,, (6)

where gμνg^{\mu\nu} is the spacetime metric, i.e., gμν=diag(1,1,1,1)g^{\mu\nu}={\rm diag}(1,-1,-1,-1) in our case. Given the coordinates wiw^{i}, we can express the ‘temporal’ direction via the normal vector

v^μ=e0μe0μe0μ,\displaystyle\hat{v}^{\mu}=\frac{e_{0}^{\mu}}{\sqrt{e_{0}^{\mu}e_{0\mu}}}\,, (7)

with

e0μ=ϵ¯μαβγe1αe2βe3γ,\displaystyle e_{0}^{\mu}=\bar{\epsilon}^{\mu\alpha\beta\gamma}\,e_{1\alpha}e_{2\beta}e_{3\gamma}\,, (8)

where |detg|ϵ¯μαβγ=sign(detg)ϵμαβγ\sqrt{|{\mathrm{det}}\,g|}\,\bar{\epsilon}^{\mu\alpha\beta\gamma}={\rm sign}({\mathrm{det}}\,g)\,\epsilon^{\mu\alpha\beta\gamma} is the covariant Levi-Civita tensor. We consider only spacelike hypersurfaces, defined by having a timelike normal vector, v^μv^μ=+1\hat{v}^{\mu}\hat{v}_{\mu}=+1.

Using these definitions, we can decompose the ambient metric as

gμν=v^μv^νΔμν,\displaystyle g^{\mu\nu}=\hat{v}^{\mu}\hat{v}^{\nu}-\Delta^{\mu\nu}\,, (9)

where Δμν=Gijeiμejν\Delta^{\mu\nu}=G^{ij}e_{i}^{\mu}e_{j}^{\nu} projects onto the hypersurface. This allows us define the temporal and spatial coordinates

x=v^μxμ,𝕩=𝕖μxμ.\displaystyle x_{\parallel}=\hat{v}^{\mu}x_{\mu}\,,\qquad\mathbb{x}_{\perp}=\mathbb{e}^{\mu}x_{\mu}\,. (10)

Thus, ‘time’ and ‘space’ are defined as projections of the ambient-space coordinates normal and tangential to Σ\Sigma. This construction defines a foliation of spacetime, where instead of Minkowski space {x0}×3\{x^{0}\}\times\mathbb{R}^{3}, we describe spacetime as {x}×Σ\{x_{\parallel}\}\times\Sigma.

Since the spatial coordinates are the components of tangent vectors xieiμx_{\perp}^{i}e_{i}^{\mu} on Σ\Sigma, the corresponding derivatives are covariant derivatives with the induced connection,

Γijk=12(Gijwk+GikwjGjkwi).\displaystyle\Gamma_{ijk}=\frac{1}{2}\Bigg{(}\frac{\partial G_{ij}}{\partial w^{k}}+\frac{\partial G_{ik}}{\partial w^{j}}-\frac{\partial G_{jk}}{\partial w^{i}}\Bigg{)}\,. (11)

This defines the intrinsic covariant derivative, which for a vector field on Σ\Sigma, ai=aμeiμa_{\perp\,i}=a_{\mu}e^{\mu}_{i}, is

jai=aiwjΓijkak.\displaystyle\partial_{\perp\,j}\,a_{\perp\,i}=\frac{\partial a_{\perp\,i}}{\partial w^{j}}-\Gamma^{k}_{ij}\,a_{\perp\,k}\,. (12)

This can also be expressed in terms of the covariant derivative of the ambient space. Since this is ordinary Minkowski space, this is just the partial derivative, and

jai=νaμeiμejν.\displaystyle\partial_{\perp\,j}\,a_{\perp\,i}=\partial_{\nu}\,a_{\mu}\,e^{\mu}_{i}e^{\nu}_{j}. (13)

Thus, for a hypersurface embedded into flat space, we can define the intrinsic covariant derivative through the projection

i=eiμμ.\displaystyle\partial_{\perp\,i}=e_{i}^{\mu}\partial_{\mu}\,. (14)

In analogy the time derivative is

=v^νν.\displaystyle\partial_{\parallel}=\hat{v}^{\nu}\partial_{\nu}\,. (15)

Note that contractions of spatial coordinates always involve the induced metric, e.g., 𝕒𝕓=Gijaibj\mathbb{a}_{\perp}\!\!\cdot\mathbb{b}_{\perp}=G_{ij}\,a_{\perp}^{i}b_{\perp}^{j}. The product of four-vectors becomes ab=ab𝕒𝕓a\!\cdot\!b=a_{\parallel}b_{\parallel}-\mathbb{a}_{\perp}\!\!\cdot\mathbb{b}_{\perp}.

Spacetime integrals translate into

d4x=dxd3w|e0μxμx|=dxd3we0μv^μdxdΣ,\displaystyle d^{4}x=dx_{\parallel}\,d^{3}w\,\bigg{|}e_{0}^{\mu}\frac{\partial x_{\mu}}{\partial x_{\parallel}}\bigg{|}=dx_{\parallel}\,d^{3}w\,e_{0}^{\mu}\hat{v}_{\mu}\equiv dx_{\parallel}\,d\Sigma\,, (16)

and we define

dΣμ=dΣv^μ.\displaystyle d\Sigma^{\mu}=d\Sigma\,\hat{v}^{\mu}\,. (17)

With the induced metric defined in Eq. (6), we can also write dΣ=|detG|d3wd\Sigma=\sqrt{|{\mathrm{det}}G|}\,d^{3}w.

We assume that the fields we are interested in can be described as quasiparticles on Σ\Sigma. The corresponding effective action of a scalar field can be written as

S=dxdΣ{12ϕ(x)[2+Z(2)2m2]ϕ(x)+J(x)ϕ(x)}.\displaystyle\begin{split}S=\int\!\!\!dx_{\parallel}\!\int\!\!\!d\Sigma\bigg{\{}&\frac{1}{2}\phi(x)\Big{[}\!-\partial_{\parallel}^{2}+Z\big{(}-\partial_{\perp}^{2}\big{)}\,\partial_{\perp}^{2}-m^{2}\Big{]}\phi(x)\\ &+J(x)\phi(x)\bigg{\}}\,.\end{split} (18)

We introduced a source J(x)J(x) which encodes interactions with external particles or fields. We allow for a general spatial ‘wave-function renormalization’ Z(2)Z\big{(}-\partial_{\perp}^{2}\big{)}, which can itself be an arbitrary function of the spatial derivative operator 2\partial_{\perp}^{2}. In a free Lorentz-invariant theory Z=1Z=1. However, if boost invariance is broken, Z1Z\neq 1 and higher-order spatial derivative terms can be induced, e.g., by spatial modulations, as in a moat regime. For example, in a low-momentum expansion up to order 4\partial_{\perp}^{4}, one recovers the effective action used in Refs. [17, 19]. Since the specific conditions where particles are in a given regime are in general met only on nontrivial hypersurfaces of, e.g., the expanding fireball of a heavy-ion collision, it is most sensible to formulate the resulting effective field theory also on this surface, instead of in flat space.

Due to the lack of Lorentz invariance, the effective action is frame-dependent. We assume Eq. (18) is the action in the local frame of the particles on Σ\Sigma, i.e., in the rest frame of the medium. Because of higher-order spatial derivative terms in the effective action, boosting into a different frame would introduce higher-order time derivatives and the resulting theory will be plagued by Ostrogradsky instabilities. We restrict ourselves to theories with two time derivatives and use ordinary canonical quantization in the following.

Given a specific foliation, we can use the well-known techniques of quantum field theory in curved spacetime to canonically quantize the theory defined by Eq. (18). To this end, we define the symplectic form

(χ1,χ2)=i𝑑Σμ[χ1(μχ2)(μχ1)χ2],\displaystyle(\chi_{1},\chi_{2})=i\int\!d\Sigma^{\mu}\big{[}\chi_{1}^{*}(\partial_{\mu}\chi_{2})-(\partial_{\mu}\chi_{1}^{*})\chi_{2}\big{]}\,, (19)

where χ1,2(x)\chi_{1,2}(x) are solutions of the equations of motion (EoM) for the action (18). We emphasize that if Lorentz invariance is broken, the symplectic form depends on the choice of the hypersurface Σ\Sigma. The free EoM,

[2Z(2)2+m2]ϕ0(x)=0,\displaystyle\Big{[}\partial_{\parallel}^{2}-Z\big{(}-\partial_{\perp}^{2}\big{)}\,\partial_{\perp}^{2}+m^{2}\Big{]}\phi_{0}(x)=0\,, (20)

are solved by plane waves

u𝕡=12ω𝕡eip¯x.\displaystyle u_{\mathbb{p}_{\perp}}=\frac{1}{\sqrt{2\omega_{\mathbb{p}_{\perp}}}}e^{-i\bar{p}\cdot x}\,. (21)

We use a notation where an overline, ¯\,\overline{\phantom{p}}\,, indicates that the momentum is on the mass-shell of the particle. Thus p¯μ\bar{p}^{\mu} is the on-shell momentum and

ω𝕡v^μp¯μ=Z(𝕡2)𝕡2+m2\displaystyle\omega_{\mathbb{p}_{\perp}}\equiv\hat{v}^{\mu}\bar{p}_{\mu}=\sqrt{Z\big{(}\mathbb{p}_{\perp}^{2}\big{)}\,\mathbb{p}_{\perp}^{2}+m^{2}} (22)

is the energy of the particle on Σ\Sigma. The waves are normalized such that

(u𝕡,u𝕢)=(2π)3δ(3)(𝕡𝕢)=𝑑Σei(𝕡𝕢)𝕩.\displaystyle(u_{\mathbb{p}_{\perp}},u_{\mathbb{q}_{\perp}})=(2\pi)^{3}\delta^{(3)}(\mathbb{p}_{\perp}-\mathbb{q}_{\perp})=\int\!d\Sigma\,e^{-i(\mathbb{p}_{\perp}-\mathbb{q}_{\perp})\cdot\mathbb{x}_{\perp}}\,. (23)

The solution ϕ0(x)\phi_{0}(x) of Eq. (20) can now be expanded in terms of these waves, with the annihilation and creation operators a~𝕡\tilde{a}_{\mathbb{p}_{\perp}} and a~𝕡\tilde{a}_{\mathbb{p}_{\perp}}^{\dagger} as coefficients,

ϕ0(x)=d3𝕡(2π)3(a~𝕡u𝕡+a~𝕡u𝕡).\displaystyle\phi_{0}(x)=\int\!\frac{d^{3}\mathbb{p}_{\perp}}{(2\pi)^{3}}\,\big{(}\tilde{a}_{\mathbb{p}_{\perp}}u_{\mathbb{p}_{\perp}}+\tilde{a}_{\mathbb{p}_{\perp}}^{\dagger}u_{\mathbb{p}_{\perp}}^{*}\big{)}\,. (24)

Using this relation, or through the symplectic form in Eq. (19), the ladder operators at time xx_{\parallel} can be expressed as

a~𝕡=(u𝕡,ϕ0(x))=i𝑑Σμeip¯x12ω𝕡(μip¯μ)ϕ0(x),a~𝕡=(u𝕡,ϕ0(x))=i𝑑Σμeip¯x12ω𝕡(μ+ip¯μ)ϕ0(x).\displaystyle\begin{split}\tilde{a}_{\mathbb{p}_{\perp}}&=\big{(}u_{\mathbb{p}_{\perp}},\phi_{0}(x)\big{)}\\ &=i\int\!d\Sigma^{\mu}\,e^{i\bar{p}\cdot x}\frac{1}{\sqrt{2\omega_{\mathbb{p}_{\perp}}}}\big{(}\partial_{\mu}-i\bar{p}_{\mu}\big{)}\phi_{0}(x)\,,\\ \tilde{a}_{\mathbb{p}_{\perp}}^{\dagger}&=-\big{(}u_{\mathbb{p}_{\perp}}^{*},\phi_{0}(x)\big{)}\\ &=-i\int\!d\Sigma^{\mu}\,e^{-i\bar{p}\cdot x}\frac{1}{\sqrt{2\omega_{\mathbb{p}_{\perp}}}}\big{(}\partial_{\mu}+i\bar{p}_{\mu}\big{)}\phi_{0}(x)\,.\end{split} (25)

This completes the quantization of the homogeneous case defined by the free EoM in Eq. (20). We are, however, interested in the more general case where particles can emerge from external fields or interactions. The corresponding EoM which follow from Eq. (18) are then

[2Z(2)2+m2]ϕ(x)=J(x).\displaystyle\Big{[}\partial_{\parallel}^{2}-Z\big{(}-\partial_{\perp}^{2}\big{)}\,\partial_{\perp}^{2}+m^{2}\Big{]}\phi(x)=J(x)\,. (26)

Under the assumption that the source JJ is only turned on for x>0x_{\parallel}>0, and by using the retarded propagator on Σ\Sigma,

DR(p)=1(p+iϵ)2+ω𝕡2,\displaystyle D_{R}(p)=\frac{1}{-(p_{\parallel}+i\epsilon)^{2}+\omega_{\mathbb{p}_{\perp}}^{2}}\,, (27)

Eq. (26) is solved by

ϕ(x)=ϕ0(x)+d4xDR(xx)J(x).\displaystyle\phi(x)=\phi_{0}(x)+\int\!d^{4}x^{\prime}\,D_{R}(x-x^{\prime})\,J(x^{\prime})\,. (28)

Consequently, the shifted ladder operators,

a𝕡=a~𝕡+i2ω𝕡d4xeip¯xJ(x),a𝕡=a~𝕡i2ω𝕡d4xeip¯xJ(x),\displaystyle\begin{split}a_{\mathbb{p}_{\perp}}&=\tilde{a}_{\mathbb{p}_{\perp}}+\frac{i}{\sqrt{2\omega_{\mathbb{p}_{\perp}}}}\int\!d^{4}x\,e^{i\bar{p}\cdot x}J(x)\,,\\ a_{\mathbb{p}_{\perp}}^{\dagger}&=\tilde{a}_{\mathbb{p}_{\perp}}^{\dagger}-\frac{i}{\sqrt{2\omega_{\mathbb{p}_{\perp}}}}\int\!d^{4}x\,e^{-i\bar{p}\cdot x}J^{*}(x)\,,\end{split} (29)

are given by

a𝕡=(u𝕡,ϕ(x))=i𝑑Σμeip¯x12ω𝕡(μip¯μ)ϕ(x),a𝕡=(u𝕡,ϕ(x))=i𝑑Σμeip¯x12ω𝕡(μ+ip¯μ)ϕ(x).\displaystyle\begin{split}a_{\mathbb{p}_{\perp}}&=\big{(}u_{\mathbb{p}_{\perp}},\phi(x)\big{)}\\ &=i\int\!d\Sigma^{\mu}\,e^{i\bar{p}\cdot x}\frac{1}{\sqrt{2\omega_{\mathbb{p}_{\perp}}}}\big{(}\partial_{\mu}-i\bar{p}_{\mu}\big{)}\phi(x)\,,\\ a_{\mathbb{p}_{\perp}}^{\dagger}&=-\big{(}u_{\mathbb{p}_{\perp}}^{*},\phi(x)\big{)}\\ &=-i\int\!d\Sigma^{\mu}\,e^{-i\bar{p}\cdot x}\frac{1}{\sqrt{2\omega_{\mathbb{p}_{\perp}}}}\big{(}\partial_{\mu}+i\bar{p}_{\mu}\big{)}\phi(x)\,.\end{split} (30)

We can use these expressions to compute any particle spectrum, such as the ones shown in Eq. (3), on an arbitrary spacelike hypersurface Σ\Sigma. Furthermore, this equation is valid for arbitrary dispersion relations ω𝕡\omega_{\mathbb{p}_{\perp}}.

For an ordinary relativistic particle, the relations in Eq. (30) are typically the starting point of the derivation of the LSZ reduction formula. There, x0±x_{0}\rightarrow\pm\infty in order to define in- and out-states. In contrast, here, xx_{\parallel} can be any time where a quasiparticle picture applies, i.e., where interactions can be considered as classical sources. In the following, we will show that this allows us to express spectra of these particles in terms of their real-time correlation functions. For applications of related ideas to the Schwinger effect and scattering amplitudes, see Refs. [37, 38].

III.2 Spectra from correlators

With the help of Eq. (30) we can express particle spectra on Σ\Sigma in terms of correlation functions of fields. Given the generating functional

Z[j]=𝒟ϕeiS[ϕ]+i𝑑x𝑑Σj(x)ϕ(x),\displaystyle Z[j]=\int\!\mathcal{D}\phi e^{iS[\phi]+i\int\!dx_{\parallel}\int\!d\Sigma\,j(x)\phi(x)}\,, (31)

we can write in general

a𝕡1a𝕡na𝕢1a𝕢n=limx,1,,x,nxy,1,,y,nx[k=1ndΣxkμdΣykν12ω𝕡keip¯kxkeiq¯kyk×(μxk+ip¯k,μ)(μykiq¯k,μ)]×(i)2nZ[0](k=n1δδj(yk)))(k=n1δδj(xk)))Z[j]|j=0,\displaystyle\begin{split}&\big{\langle}a_{\mathbb{p}_{1}}^{\dagger}\cdots a_{\mathbb{p}_{n}}^{\dagger}a_{\mathbb{q}_{1}}\cdots a_{\mathbb{q}_{n}}\big{\rangle}\\ &=\lim_{\begin{subarray}{c}x_{\parallel,1},\dots,x_{\parallel,n}\rightarrow x_{\parallel}\\ y_{\parallel,1},\dots,y_{\parallel,n}\rightarrow x_{\parallel}\end{subarray}}\,\Bigg{[}\prod_{k=1}^{n}\int\!d\Sigma_{x_{k}}^{\mu}\,d\Sigma_{y_{k}}^{\nu}\,\frac{1}{2\omega_{\mathbb{p}_{k}}}\,e^{-i\bar{p}_{k}\cdot x_{k}}e^{i\bar{q}_{k}\cdot y_{k}}\\ &\quad\times\big{(}\partial_{\mu}^{x_{k}}+i\bar{p}_{k,\mu}\big{)}\big{(}\partial_{\mu}^{y_{k}}-i\bar{q}_{k,\mu}\big{)}\Bigg{]}\\ &\quad\times\frac{(-i)^{2n}}{Z[0]}\Bigg{(}\prod_{k=n}^{1}\frac{\delta}{\delta j(y_{k}))}\Bigg{)}\Bigg{(}\prod_{k=n}^{1}\frac{\delta}{\delta j(x_{k}))}\Bigg{)}\,Z[j]\Bigg{|}_{j=0}\,,\end{split} (32)

Note that the path integral automatically generates time-ordered correlation functions. The above expression only describes spectra on Σ\Sigma after the equal-time limit is taken. This is indicated by the limit in Eq. (32), where all times x,kx_{\parallel,k} and y,ky_{\parallel,k} are set to the same time xx_{\parallel}. In the following, we explicitly work out the expressions for single-particle spectra. Since this corresponds to two-point functions of the fields, this is sufficient to describe all nn-particle spectra if the generating functional is Gaussian. All correlations/spectra can then be expressed in terms of two-point functions/single-particle spectra by virtue of Wick’s theorem.

For illustration, and since it is most relevant to the application below, see Sec. IV.1, we compute the mixed single-particle spectrum,

n1(𝕡,𝕢)=ω𝕡ω𝕢a𝕡a𝕢.\displaystyle n_{1}(\mathbb{p}_{\perp},\mathbb{q}_{\perp})=\sqrt{\omega_{\mathbb{p}_{\perp}}\omega_{\mathbb{q}_{\perp}}}\,\big{\langle}a_{\mathbb{p}_{\perp}}^{\dagger}a_{\mathbb{q}_{\perp}}\big{\rangle}\,. (33)

For 𝕡=𝕢\mathbb{p}_{\perp}=\mathbb{q}_{\perp} this reduces to the ordinary single-particle spectrum in Eq. (3). Using Eq. (30) we get

n1(𝕡,𝕢)=12limyx𝑑Σxμ𝑑Σyνeip¯xeiq¯y×(μx+ip¯μ)(νyiq¯ν)ϕ(x)ϕ(y).\displaystyle\begin{split}n_{1}(\mathbb{p}_{\perp},\mathbb{q}_{\perp})&=\frac{1}{2}\lim_{y_{\parallel}\rightarrow x_{\parallel}}\int\!d\Sigma_{x}^{\mu}\,d\Sigma_{y}^{\nu}\,e^{-i\bar{p}\cdot x}e^{i\bar{q}\cdot y}\\ &\quad\times\big{(}\partial^{x}_{\mu}+i\bar{p}_{\mu}\big{)}\big{(}\partial^{y}_{\nu}-i\bar{q}_{\nu}\big{)}\langle\phi(x)\phi(y)\rangle\,.\end{split} (34)

Since the particle spectrum is defined at equal time, so is the correlation function on the right-hand side. In order to express this as something resembling a phase-space distribution, we define the average and relative positions

X=12(x+y),ΔX=xy.\displaystyle X=\frac{1}{2}(x+y)\,,\qquad\Delta X=x-y\,. (35)

By also introducing the average and relative momenta,

P=12(p+q),ΔP=pq,P¯=12(p¯+q¯),ΔP¯=p¯q¯,\displaystyle\begin{split}P&=\frac{1}{2}(p+q)\,,\qquad\Delta P=p-q\,,\\ \overline{P}&=\frac{1}{2}(\bar{p}+\bar{q})\,,\qquad\overline{\Delta P}=\bar{p}-\bar{q}\,,\end{split} (36)

the mixed spectrum becomes

n1(𝕡,𝕢)=12limΔX0𝑑ΣXμ𝑑ΣΔXνeiΔP¯XeiP¯ΔX×~μνϕ(X+12ΔX)ϕ(X12ΔX).\displaystyle\begin{split}n_{1}(\mathbb{p}_{\perp},\mathbb{q}_{\perp})&=\frac{1}{2}\lim_{\Delta X_{\parallel}\rightarrow 0}\int\!d\Sigma_{X}^{\mu}\,d\Sigma_{\Delta X}^{\nu}\,e^{-i\overline{\Delta P}\cdot X}e^{-i\overline{P}\cdot\Delta X}\\ &\quad\times\widetilde{\Box}_{\mu\nu}\,\bigg{\langle}\phi\bigg{(}X+\frac{1}{2}\Delta X\bigg{)}\phi\bigg{(}X-\frac{1}{2}\Delta X\bigg{)}\bigg{\rangle}\,.\end{split} (37)

Here, dΣXμ=dΣXv^μd\Sigma_{X}^{\mu}=d\Sigma_{X}\hat{v}^{\mu} and dΣΔXν=dΣΔXv^νd\Sigma_{\Delta X}^{\nu}=d\Sigma_{\Delta X}\hat{v}^{\nu} are the integration measures for the average and relative hypersurfaces at times X=v^μXμX_{\parallel}=\hat{v}^{\mu}X_{\mu} and ΔX=v^μΔXμ\Delta X_{\parallel}=\hat{v}^{\mu}\Delta X_{\mu}. The operator ~μν\widetilde{\Box}_{\mu\nu} is

~μν=14μXνX12μXνΔX+12μΔXνXμΔXνΔXi2(P¯ν12ΔP¯ν)(μX+2μΔX)+i2(P¯μ+12ΔP¯μ)(νX2νΔX)+(P¯μ+12ΔP¯μ)(P¯ν12ΔP¯ν).\displaystyle\begin{split}\widetilde{\Box}_{\mu\nu}&=\frac{1}{4}\partial_{\mu}^{X}\partial_{\nu}^{X}-\frac{1}{2}\partial_{\mu}^{X}\partial_{\nu}^{\Delta X}+\frac{1}{2}\partial_{\mu}^{\Delta X}\partial_{\nu}^{X}-\partial_{\mu}^{\Delta X}\partial_{\nu}^{\Delta X}\\ &\quad-\frac{i}{2}\bigg{(}\overline{P}_{\nu}-\frac{1}{2}\overline{\Delta P}_{\nu}\bigg{)}\big{(}\partial_{\mu}^{X}+2\partial_{\mu}^{\Delta X}\big{)}\\ &\quad+\frac{i}{2}\bigg{(}\overline{P}_{\mu}+\frac{1}{2}\overline{\Delta P}_{\mu}\bigg{)}\big{(}\partial_{\nu}^{X}-2\partial_{\nu}^{\Delta X}\big{)}\\ &\quad+\bigg{(}\overline{P}_{\mu}+\frac{1}{2}\overline{\Delta P}_{\mu}\bigg{)}\bigg{(}\overline{P}_{\nu}-\frac{1}{2}\overline{\Delta P}_{\nu}\bigg{)}\,.\end{split} (38)

Through the integrations over dΣXμd\Sigma_{X}^{\mu} and dΣΔXμd\Sigma_{\Delta X}^{\mu} this operator is projected normal to these hypersurfaces,

v^μv^ν~μν=14(X)2(ΔX)2+i2ΔP¯X2iP¯ΔX+P¯214ΔP¯2,\displaystyle\begin{split}\hat{v}^{\mu}\hat{v}^{\nu}\widetilde{\Box}_{\mu\nu}&=\frac{1}{4}\big{(}\partial_{\parallel}^{X}\big{)}^{2}-\big{(}\partial_{\parallel}^{\Delta X}\big{)}^{2}+\frac{i}{2}\overline{\Delta P}_{\parallel}\partial_{\parallel}^{X}-2i\overline{P}_{\parallel}\partial_{\parallel}^{\Delta X}\\ &\quad+\overline{P}_{\parallel}^{2}-\frac{1}{4}\overline{\Delta P}_{\parallel}^{2}\,,\end{split} (39)

where P¯=12(ω𝕡+ω𝕢)\overline{P}_{\parallel}=\frac{1}{2}(\omega_{\mathbb{p}_{\perp}}+\omega_{\mathbb{q}_{\perp}}) and ΔP¯=ω𝕡ω𝕢\overline{\Delta P}_{\parallel}=\omega_{\mathbb{p}_{\perp}}-\omega_{\mathbb{q}_{\perp}}. The integration over ΔX\Delta X in Eq. (37) is almost a Wigner transformation, which replaces the relative position ΔX\Delta X by the average momentum P¯\overline{P}. What is missing is the ΔX\Delta X_{\parallel} integration. It can be included by introducing and then cancelling the corresponding Fourier transformation (note that this is a transformation with negative momentum P-P),

h(ΔX)=dP2πeiPΔX𝑑ΔXeiPΔXh(ΔX),\displaystyle h(\Delta X_{\parallel})=\int\!\frac{dP_{\parallel}}{2\pi}\,e^{iP_{\parallel}\Delta X_{\parallel}}\int\!d\Delta X_{\parallel}^{\prime}\,e^{-iP_{\parallel}\Delta X_{\parallel}^{\prime}}\,h(\Delta X_{\parallel}^{\prime})\,, (40)

where hh is an arbitrary function. Applying this to Eq. (37), we can carry out the full Fourier transform with respect to ΔX\Delta X. We define the Wigner-transformed spectral and statistical functions from the commutators and the anti-commutators of the fields,

ρ(X,P)=𝑑ΔX𝑑ΣΔXeiPΔX×[ϕ(X+12ΔX),ϕ(X12ΔX)],F(X,P)=12𝑑ΔX𝑑ΣΔXeiPΔX×{ϕ(X+12ΔX),ϕ(X12ΔX)}.\displaystyle\begin{split}\rho(X,P)&=\int\!d\Delta X_{\parallel}\int\!d\Sigma_{\Delta X}\,e^{iP\cdot\Delta X}\\ &\quad\times\bigg{\langle}\bigg{[}\phi\bigg{(}X+\frac{1}{2}\Delta X\bigg{)},\phi\bigg{(}X-\frac{1}{2}\Delta X\bigg{)}\bigg{]}\bigg{\rangle}\,,\\ F(X,P)&=\frac{1}{2}\int\!d\Delta X_{\parallel}\int\!d\Sigma_{\Delta X}\,e^{iP\cdot\Delta X}\\ &\quad\times\bigg{\langle}\bigg{\{}\phi\bigg{(}X+\frac{1}{2}\Delta X\bigg{)},\phi\bigg{(}X-\frac{1}{2}\Delta X\bigg{)}\bigg{\}}\bigg{\rangle}\,.\end{split} (41)

With this, and taking the limit ΔX0\Delta X_{\parallel}\rightarrow 0, we finally arrive at

n1(𝕡,𝕢)=12𝑑ΣXeiΔP¯XdP2π×[14(X)2+i2ΔP¯X+(P+P¯)214ΔP¯2]×[F(X,P)12ρ(X,P)].\displaystyle\begin{split}n_{1}(\mathbb{p}_{\perp},\mathbb{q}_{\perp})&=\frac{1}{2}\int\!d\Sigma_{X}\,e^{-i\overline{\Delta P}\cdot X}\int\!\frac{dP_{\parallel}}{2\pi}\\ &\quad\times\bigg{[}\frac{1}{4}\big{(}\partial_{\parallel}^{X}\big{)}^{2}+\frac{i}{2}\overline{\Delta P}_{\parallel}\partial_{\parallel}^{X}+\big{(}P_{\parallel}+\overline{P}_{\parallel}\big{)}^{2}-\frac{1}{4}\overline{\Delta P}_{\parallel}^{2}\bigg{]}\\ &\quad\times\bigg{[}F(X,P)-\frac{1}{2}\rho(X,P)\bigg{]}\,.\end{split} (42)

Note that it was important not to go to equal times from the beginning. Otherwise we would not have been able to do the Wigner transformation.

Equation (42) is the most general expression we can derive for the mixed particle spectrum. We see that the relative momentum ΔP¯\overline{\Delta P} is correlated with the average location XX. The Wigner-transformed correlation functions are functions of the average pair momentum PP.

To get the single-particle spectrum, we set p=qp=q, so that ΔP¯=0\overline{\Delta P}=0 and P=pP=p, and

n1(𝕡)=12𝑑ΣXdp2π[14(X)2+(p+ω𝕡)2]×[F(X,p)12ρ(X,p)],\displaystyle\begin{split}n_{1}(\mathbb{p}_{\perp})&=\frac{1}{2}\int\!d\Sigma_{X}\,\int\!\frac{dp_{\parallel}}{2\pi}\,\bigg{[}\frac{1}{4}\big{(}\partial_{\parallel}^{X}\big{)}^{2}+\big{(}p_{\parallel}+\omega_{\mathbb{p}_{\perp}}\big{)}^{2}\bigg{]}\\ &\quad\times\bigg{[}F(X,p)-\frac{1}{2}\rho(X,p)\bigg{]}\,,\end{split} (43)

since the on-shell longitudinal momentum is p¯=ω𝕡\bar{p}_{\parallel}=\omega_{\mathbb{p}_{\perp}}. Although these expressions have been derived assuming a quasiparticle picture, they can in principle be used to define the spectra non-perturbatively and out of equilibrium since they depend on the general spectral and statistical functions of the theory.

We can simplify these expressions by making a few assumptions which are reasonable if one has, as we do, the hydrodynamic regime of a heavy-ion collision in mind. First, we assume that the system is in local thermodynamic equilibrium. This means that FF and ρ\rho are related through a generalized fluctuation-dissipation relation (sometimes called the Kadanoff-Baym ansatz) [39],

F(X,P)=[12+f(X,P)]ρ(X,P),\displaystyle F(X,P)=\bigg{[}\frac{1}{2}+f(X,P)\bigg{]}\rho(X,P)\,, (44)

where f(X,P)f(X,P) is the single-particle distribution. Using this relation, the Boltzmann equation for f(X,P)f(X,P) can be derived from the Kadanoff-Baym equations [40].

Furthermore, we will neglect any derivatives in XX. This is certainly true for isotropic systems in thermodynamic equilibrium, where only relative positions matter. In general, such gradients are accompanied by additional powers of \hbar and can be neglected in a semi-classical approximation.

Using these approximations, we finally arrive at

n1(𝕡,𝕢)=12𝑑ΣXeiΔP¯XdP2π[(P+P¯)214ΔP¯2]×f(X,P)ρ(X,P),n1(𝕡)=12𝑑ΣXdp2π(p+ω𝕡)2f(X,p)ρ(X,p).\displaystyle\begin{split}n_{1}(\mathbb{p}_{\perp},\mathbb{q}_{\perp})&=\frac{1}{2}\int\!d\Sigma_{X}\,e^{-i\overline{\Delta P}\cdot X}\int\!\frac{dP_{\parallel}}{2\pi}\bigg{[}\big{(}P_{\parallel}+\overline{P}_{\parallel}\big{)}^{2}-\frac{1}{4}\overline{\Delta P}_{\parallel}^{2}\bigg{]}\\ &\quad\times f(X,P)\,\rho(X,P)\,,\\ n_{1}(\mathbb{p}_{\perp})&=\frac{1}{2}\int\!d\Sigma_{X}\int\!\frac{dp_{\parallel}}{2\pi}\,\big{(}p_{\parallel}+\omega_{\mathbb{p}_{\perp}}\big{)}^{2}\,f(X,p)\,\rho(X,p)\,.\end{split} (45)

These equations can be viewed as generalized Cooper-Frye formulas [32]. In fact, if we use a covariant Bose-Einstein distribution for ff and the spectral function of a Lorentz-invariant free particle, n1(𝕡)n_{1}(\mathbb{p}_{\perp}) reduces to the original expression of Cooper and Frye, see Sec. IV.2. Thus, with our formalism we have derived the phenomenological Cooper-Frye formula from the underlying microscopic quantum field theory. Equation (45) also explicitly accounts for off-shell effects. The advantage of our formalism is its generality. The original versions of the Cooper-Frye formalism and HBT, which have been derived in a similar manner as in the present work in Ref. [33], are limited to free particles on the mass-shell. We are not restricted to these limitations here.

IV HBT on a moat hypersurface

IV.1 HBT in the Gaussian approximation

With the procedure from above any nn-particle spectrum can be expressed in terms of 2n2n-point functions of fields. For simplicity, we assume that the generating functional in Eq. (31) is Gaussian. As mentioned above, in the Gaussian approximation nn-particle spectra can be expressed solely in terms of single-particle ones. This turns out to be sufficient for the theoretical description of HBT correlations. Here we explicitly work out the two-particle spectrum n2(𝕡,𝕢)n_{2}(\mathbb{p}_{\perp},\mathbb{q}_{\perp}), defined in Eq. (3). In the Gaussian approximation the 4-point function is

ϕ(x1)ϕ(x2)ϕ(y1)ϕ(y2)=ϕ(x1)ϕ(y1)ϕ(x2)ϕ(y2)+ϕ(x1)ϕ(y2)ϕ(x2)ϕ(y1)+ϕ(x1)ϕ(x2)ϕ(y1)ϕ(y2).\displaystyle\begin{split}&\big{\langle}\phi(x_{1})\phi(x_{2})\phi(y_{1})\phi(y_{2})\big{\rangle}\\ &\qquad=\big{\langle}\phi(x_{1})\phi(y_{1})\big{\rangle}\big{\langle}\phi(x_{2})\phi(y_{2})\big{\rangle}\\ &\qquad\quad+\big{\langle}\phi(x_{1})\phi(y_{2})\big{\rangle}\big{\langle}\phi(x_{2})\phi(y_{1})\big{\rangle}\\ &\qquad\quad+\big{\langle}\phi(x_{1})\phi(x_{2})\big{\rangle}\big{\langle}\phi(y_{1})\phi(y_{2})\big{\rangle}\,.\end{split} (46)

By comparison with Eq. (32), we can identify the different contributions:

(a)ϕ(x1)ϕ(y1)ϕ(x2)ϕ(y2)n1(𝕡)n1(𝕢),(b)ϕ(x1)ϕ(y2)ϕ(x2)ϕ(y1)n1(𝕡,𝕢)n1(𝕢,𝕡)=|n1(𝕡,𝕢)|2,(c)ϕ(x1)ϕ(x2)ϕ(y1)ϕ(y2)ω𝕡ω𝕢a𝕡a𝕢a𝕡a𝕢|n¯1(𝕡,𝕢)|2.\displaystyle\begin{split}&{\rm(a)}\;\;\big{\langle}\phi(x_{1})\phi(y_{1})\big{\rangle}\big{\langle}\phi(x_{2})\phi(y_{2})\big{\rangle}\\ &\qquad\longrightarrow\,n_{1}(\mathbb{p}_{\perp})\,n_{1}(\mathbb{q}_{\perp})\,,\\[5.0pt] &{\rm(b)}\;\;\big{\langle}\phi(x_{1})\phi(y_{2})\big{\rangle}\big{\langle}\phi(x_{2})\phi(y_{1})\big{\rangle}\\ &\qquad\longrightarrow\,n_{1}(\mathbb{p}_{\perp},\mathbb{q}_{\perp})\,n_{1}(\mathbb{q}_{\perp},\mathbb{p}_{\perp})=\big{|}n_{1}(\mathbb{p}_{\perp},\mathbb{q}_{\perp})\big{|}^{2}\,,\\[5.0pt] &{\rm(c)}\;\;\big{\langle}\phi(x_{1})\phi(x_{2})\big{\rangle}\big{\langle}\phi(y_{1})\phi(y_{2})\big{\rangle}\\ &\qquad\longrightarrow\omega_{\mathbb{p}_{\perp}}\omega_{\mathbb{q}_{\perp}}\,\big{\langle}a_{\mathbb{p}_{\perp}}^{\dagger}a_{\mathbb{q}_{\perp}}^{\dagger}\big{\rangle}\big{\langle}a_{\mathbb{p}_{\perp}}a_{\mathbb{q}_{\perp}}\big{\rangle}\equiv\big{|}\bar{n}_{1}(\mathbb{p}_{\perp},\mathbb{q}_{\perp})\big{|}^{2}\,.\end{split} (47)

(a) and (b) we already computed. (a) is just a product of two single-particle distributions. (b) stems from particle-particle interference. This is the actual HBT effect. It describes the correlation of identical particles generated by their Bose-Einstein statistics. (c) can be interpreted as particle-antiparticle interference. For particles that are their own antiparticles, like the neutral scalar fields considered here, we cannot distinguish this. Otherwise, one of the ladder operators in each expectation value would correspond to the creation/annihilation operator of an antiparticle. This is what is called the ‘surprising effect’ in Ref. [27].

The resulting two-particle spectrum is

n2(𝕡,𝕢)=n1(𝕡)n1(𝕢)+|n1(𝕡,𝕢)|2+|n¯1(𝕡,𝕢)|2.\displaystyle\begin{split}n_{2}(\mathbb{p}_{\perp},\mathbb{q}_{\perp})&=n_{1}(\mathbb{p}_{\perp})\,n_{1}(\mathbb{q}_{\perp})+\big{|}n_{1}(\mathbb{p}_{\perp},\mathbb{q}_{\perp})\big{|}^{2}\\ &\quad+\big{|}\bar{n}_{1}(\mathbb{p}_{\perp},\mathbb{q}_{\perp})\big{|}^{2}\,.\end{split} (48)

The first two terms are usually considered when studying HBT correlations. Once the spectral function, the single-particle distribution function, and the hypersurface are specified, we can compute this using Eq. (45). The missing particle-antiparticle interference term is computed in App. A. In the following, we set up appropriate models for the required quantities in order to illustrate the relevant qualitative effects.

But first, we note that HBT is often expressed in terms of the emission function (or source function) S(x,)S(x,\mathbb{P}_{\perp}), where xx is a spacetime coordinate. It describes the phase-space distribution, i.e., the distribution of spacetime positions and momenta of the particles emitted from a phase-space element d3xd3Pd^{3}x\,d^{3}P_{\perp}. The particle-particle interference term is then given by the (on-shell) Fourier transformation of the emission function with respect to the relative momentum [24, 26],

n1(𝕡,𝕢)d4xeiΔP¯xS(x,).\displaystyle n_{1}(\mathbb{p}_{\perp},\mathbb{q}_{\perp})\approx\int\!d^{4}x\,e^{-i\overline{\Delta P}\cdot x}\,S(x,\mathbb{P}_{\perp})\,. (49)

In this form, the connection between the behavior of the correlation function and the spacetime structure of the emitting source is most transparent: the range of the correlation in relative momentum Δ\mathbb{\Delta P}_{\perp} is related to the inverse size of the source. More precisely, it relates to the region of homogeneity inside the source for a given average momentum \mathbb{P}_{\perp} [41].

An underlying assumption in Eq. (49) is that the dependence of the correlation on the relative momentum Δ\mathbb{\Delta P}_{\perp} arises solely from the Fourier transformation of the location xx. As is evident from Eq. (42), this is not the case here. Instead, we can define the emission function plus a correction, S(x,)+ΔS(x,,Δ)S(x,\mathbb{P}_{\perp})+\Delta S(x,\mathbb{P}_{\perp},\mathbb{\Delta P}_{\perp}), with the emission function

S(x,)=12𝑑ΣXδ(Xx)dP2π[14(X)2+(P+P¯)2]×[F(X,P)12ρ(X,P)]12𝑑ΣXδ(Xx)dP2π(P+P¯)2f(X,P)ρ(X,P),\displaystyle\begin{split}&S(x,\mathbb{P}_{\perp})\\ &\quad=\frac{1}{2}\int\!d\Sigma_{X}\,\delta(X-x)\int\!\frac{dP_{\parallel}}{2\pi}\,\bigg{[}\frac{1}{4}\big{(}\partial^{X}_{\parallel}\big{)}^{2}+\big{(}P_{\parallel}+\overline{P}_{\parallel}\big{)}^{2}\bigg{]}\\ &\quad\quad\times\bigg{[}F(X,P)-\frac{1}{2}\rho(X,P)\bigg{]}\\ &\quad\approx\frac{1}{2}\int\!d\Sigma_{X}\,\delta(X-x)\int\!\frac{dP_{\parallel}}{2\pi}\,\big{(}P_{\parallel}+\overline{P}_{\parallel}\big{)}^{2}f(X,P)\,\rho(X,P)\,,\end{split} (50)

and the correction

ΔS(x,,Δ)=14𝑑ΣXδ(Xx)dP2π(iΔP¯X12ΔP¯2)×[F(X,P)12ρ(X,P)]18ΔP¯2𝑑ΣXδ(Xx)dP2πf(X,P)ρ(X,P).\displaystyle\begin{split}&\Delta S(x,\mathbb{P}_{\perp},\mathbb{\Delta P}_{\perp})\\ &\quad=\frac{1}{4}\int\!d\Sigma_{X}\,\delta(X-x)\int\!\frac{dP_{\parallel}}{2\pi}\,\bigg{(}i\,\overline{\Delta P}_{\parallel}\partial^{X}_{\parallel}-\frac{1}{2}\overline{\Delta P}_{\parallel}^{2}\bigg{)}\\ &\quad\quad\times\bigg{[}F(X,P)-\frac{1}{2}\rho(X,P)\bigg{]}\\ &\quad\approx-\frac{1}{8}\,\overline{\Delta P}_{\parallel}^{2}\int\!d\Sigma_{X}\,\delta(X-x)\int\!\frac{dP_{\parallel}}{2\pi}\,f(X,P)\,\rho(X,P)\,.\end{split} (51)

In the last lines of both equations we again assumed local thermodynamic equilibrium and neglected gradients in XX. With this we have

n1(𝕡,𝕢)=d4xeiΔP¯x[S(x,)+ΔS(x,,Δ)].\displaystyle n_{1}(\mathbb{p}_{\perp},\mathbb{q}_{\perp})=\int\!d^{4}x\,e^{-i\overline{\Delta P}\cdot x}\,\big{[}S(x,\mathbb{P}_{\perp})+\Delta S(x,\mathbb{P}_{\perp},\mathbb{\Delta P}_{\perp})\big{]}\,. (52)

Thus, the assumption underlying Eq. (49) that the HBT correlation can be described as the Fourier transformation of a phase-space distribution is only warranted if the difference of the energies of the on-shell particles on the hypersurface, ΔP¯\overline{\Delta P}_{\parallel}, can be neglected. This is obviously not true in general. We will discuss the validity of neglecting this contribution in Sec. IV.4.

IV.2 Moat quasiparticle

For the spectral function that enters Eq. (45) we use a similar approximation to the one used in Ref. [19], where thermodynamic fluctuations in a moat regime have been studied. For consistency with our derivation in Sec. III, we assume a quasiparticle as described by the action in Eq. (18). The retarded propagator on Σ\Sigma is then given by Eq. (27), which results in the spectral function

ρ(p)=2ImDR(p)=πω𝕡[δ(pω𝕡)δ(p+ω𝕡)].\displaystyle\begin{split}\rho(p)&=2\,{\rm Im}D_{R}(p)\\ &=\frac{\pi}{\omega_{\mathbb{p}_{\perp}}}\,\big{[}\delta(p_{\parallel}-\omega_{\mathbb{p}_{\perp}})-\delta(p_{\parallel}+\omega_{\mathbb{p}_{\perp}})\big{]}\,.\end{split} (53)

In general, any spectral function with a sufficiently sharp peak is suitable for our formalism. We choose a quasiparticle with infinite lifetime for simplicity.

For a quasiparticle with energy pp_{\parallel} on Σ\Sigma, the relativistic Boltzmann equation in local thermodynamic equilibrium and in absence of external forces yields the single-particle distribution function

f(X,p)=θ(p)nB(p)=θ(p)ep/T1.\displaystyle f(X,p)=\theta(p_{\parallel})\,n_{B}(p_{\parallel})=\frac{\theta(p_{\parallel})}{e^{p_{\parallel}/T}-1}\,. (54)

ff is a number density in phase space and as such has to be non-negative. This is ensured by the theta function.

The spectral and distribution functions depend on the location of the particle in spacetime through the projection of the momentum onto the hypersurface. Plugging Eqs. (53) and (54) into Eqs. (45) and (72) yields for the single-particle spectrum,

n1(𝕡)=𝑑ΣXω𝕡nB(ω𝕡),\displaystyle\begin{split}n_{1}(\mathbb{p}_{\perp})=\int\!d\Sigma_{X}\,\omega_{\mathbb{p}_{\perp}}\,n_{B}(\omega_{\mathbb{p}_{\perp}})\,,\end{split} (55)

for the particle-particle interference contribution,

n1(𝕡,𝕢)=14dΣX{eiΔP¯X1ω[(ω+P¯)214ΔP¯2]×nB(ω)},\displaystyle\begin{split}&n_{1}(\mathbb{p}_{\perp},\mathbb{q}_{\perp})=\\ &\quad\frac{1}{4}\int\!d\Sigma_{X}\Bigg{\{}e^{-i\overline{\Delta P}\cdot X}\frac{1}{\omega_{\mathbb{P}_{\perp}}}\bigg{[}\big{(}\omega_{\mathbb{P}_{\perp}}+\overline{P}_{\parallel}\big{)}^{2}-\frac{1}{4}\overline{\Delta P}_{\parallel}^{2}\bigg{]}\\ &\quad\times n_{B}(\omega_{\mathbb{P}_{\perp}})\Bigg{\}}\,,\end{split} (56)

and for the particle-antiparticle interference term,

n¯1(𝕡,𝕢)=18dΣX{e2iP¯X1ωΔ/2[(2ωΔ/2+ΔP¯)24P¯2]nB(ωΔ/2)}.\displaystyle\begin{split}&\bar{n}_{1}(\mathbb{p}_{\perp},\mathbb{q}_{\perp})=\\ &\quad-\frac{1}{8}\int\!d\Sigma_{X}\Bigg{\{}e^{-2i\overline{P}\cdot X}\frac{1}{\omega_{\mathbb{\Delta P}_{\perp}/2}}\bigg{[}\big{(}2\omega_{\mathbb{\Delta P}_{\perp}/2}+\overline{\Delta P}_{\parallel}\big{)}^{2}\\ &\quad-4\overline{P}_{\parallel}^{2}\bigg{]}\,n_{B}(\omega_{\mathbb{\Delta P}_{\perp}/2})\Bigg{\}}\,.\end{split} (57)

We see that with the present approximations, we recover the original Cooper-Frye formula for the single-particle spectrum n1(𝕡)n_{1}(\mathbb{p}_{\perp}), but with a more general dispersion relation. We have therefore provided a microscopic derivation for the expression proposed in Ref. [19].

The interference terms (56) and (57) receive contributions from both on- and off-shell momenta of the individual particles. In the quasiparticle picture also the on-shell average and relative energies ω\omega_{\mathbb{P}_{\perp}} and ωΔ/2\omega_{\mathbb{\Delta P}_{\perp}/2} contribute. They are manifestly different from the average and relative energies of the on-shell particles, P¯\overline{P}_{\parallel} and ΔP¯\overline{\Delta P}_{\parallel}. This reflects the generality of our formalism, as it is able to capture off-shell effects by taking into account the full spectrum of the particles involved. In previous work it was necessary to resort to on-shell approximations to describe interference [24].

Furthermore, our microscopic derivation reveals a different momentum dependence of the particle-particle interference contribution (56) as compared to the literature, cf., e.g., Refs. [42, 24]. In previous work where correlations on a sharp hypersurface have been considered (as is customary in hydrodynamical applications), it has always been assumed that interference can be described by the ordinary Cooper-Frye formula for average momenta, and an additional Fourier transformation with respect to the average position of the particles 111See, e.g., Eq. (85) in [24]. Our results show that this is not correct, as the particles on the hypersurface cannot be described by their average momentum alone.

To be specific, in addition to the assumptions that lead from the most general Eqs. (42), (43), and (71) to the above equations, our results reduce to the ones in the literature [24, 26] using the following assumptions: First, the contribution from particle-antiparticle interference, n¯1(𝕡,𝕢)\bar{n}_{1}(\mathbb{p}_{\perp},\mathbb{q}_{\perp}) needs to be neglected. Second, the difference between the on-shell average energy ω\omega_{\mathbb{P}_{\perp}} and the average of the on-shell energies P¯\overline{P}_{\parallel} has to be neglected, P¯=ω\overline{P}_{\parallel}=\omega_{\mathbb{P}_{\perp}}. Furthermore, the relative on-shell energy needs to be neglected as well, ΔP¯=0\overline{\Delta P}_{\parallel}=0. Evidently, these assumptions cannot be true in general, so the known equations in the literature have to be considered approximations of the more general ones derived here. Since a quantitative analysis is not in the scope of this work, we defer an analysis of the validity of these approximations to future work.

Refer to caption
Figure 4: Energy as a function of momentum. The gray lines are in the normal phase and the yellow lines are in a moat regime. zz is the coefficient of the momentum-squared term of the energy. In the normal phase it is E𝕡=z𝕡𝟚+m2E_{\mathbb{p}}=\sqrt{z\,\mathbb{p^{2}}+m^{2}} with z>0z>0. In the moat regime, the energy is defined in Eq. (58), where z=1λ2M2z=1-\frac{\lambda^{2}}{M^{2}}. Negative zz signals a moat regime.

IV.3 A simple model

We want to apply the formalism developed here to particles in a moat regime. To this end, we need to specify the wave-function renormalization Z(𝕡2)Z\big{(}\mathbb{p}_{\perp}^{2}\big{)}, or, equivalently, the dispersion relation (22). For simplicity, we specify ZZ in ambient-space coordinates and then project the resulting on-shell momentum normal to Σ\Sigma. Our ansatz is based on two requirements. First, the resulting energy ω𝕡\omega_{\mathbb{p}_{\perp}} should have a minimum at nonzero momentum and, second, it should have a well-defined large-momentum limit. The latter requirement entails that the energy grows linearly with momentum for large momenta, as otherwise causality would be violated. Motivated by Ref. [44], we choose

E𝕡2=Z(𝕡2)𝕡2+m2=(1λ2𝕡2+M2)𝕡2+m2(1λ2M2)𝕡2+λ2M4𝕡4+𝒪(𝕡6),\displaystyle\begin{split}E_{\mathbb{p}}^{2}&=Z\big{(}\mathbb{p}^{2}\big{)}\,\mathbb{p}^{2}+m^{2}\\ &=\bigg{(}1-\frac{\lambda^{2}}{\mathbb{p}^{2}+M^{2}}\bigg{)}\,\mathbb{p}^{2}+m^{2}\\ &\approx\bigg{(}1-\frac{\lambda^{2}}{M^{2}}\bigg{)}\,\mathbb{p}^{2}+\frac{\lambda^{2}}{M^{4}}\mathbb{p}^{4}+\mathcal{O}\big{(}\mathbb{p}^{6}\big{)}\,,\end{split} (58)

where in the second line we did a low-momentum expansion. Thus, for small momenta we have the energy E𝕡z𝕡+w𝕡4+m2E_{\mathbb{p}}\approx\sqrt{z\,\mathbb{p}+w\,\mathbb{p}^{4}+m^{2}} mentioned in Eq. (1), where z=1λ2M2z=1-\frac{\lambda^{2}}{M^{2}} and w=λ2M4w=\frac{\lambda^{2}}{M^{4}}. Such a dispersion relation has been assumed in Refs. [17, 19]. The energy has a negative 𝕡2\mathbb{p}^{2}-coefficient if λ2M2>0\frac{\lambda^{2}}{M^{2}}>0, and the system is in a moat regime. In this case, the energy is minimized at a nonzero momentum, |𝕡min|=M(λM)|\mathbb{p}_{\rm min}|=\sqrt{M(\lambda-M)}, with minimal energy Emin=m2(λM)2E_{\rm min}=\sqrt{m^{2}-(\lambda-M)^{2}}. This is illustrated in Fig. 4.

Equation (58) yields the on-shell momentum p¯μ=(E𝕡,𝕡)\bar{p}^{\mu}=(E_{\mathbb{p}},\mathbb{p}), and the resulting energy on the hypersurface is ω𝕡=v^μp¯μ\omega_{\mathbb{p}_{\perp}}=\hat{v}^{\mu}\bar{p}_{\mu}. The advantage of our ansatz for ZZ in Eq. (58) is that the corresponding particles still behave as free relativistic ones at large momenta, 𝕡2M2\mathbb{p}^{2}\gtrsim M^{2}.

It is left to specify the hypersurface ΣX\Sigma_{X}. In general it can be extracted from hydrodynamical simulations. To illustrate the basic physics we use a simple model.

We start by setting up the coordinates we use and the parametrization of ΣX\Sigma_{X}. We assume that the hypersurface ΣX\Sigma_{X} is defined by a constant proper time, τ=τΣ\tau=\tau_{\Sigma}, in ‘beam-direction’ X3X_{3}, where τ=X02X32\tau=\sqrt{X_{0}^{2}-X_{3}^{2}} and XμX_{\mu} are the Cartesian coordinates of the ambient Minkowski space. This is a common and phenomenologically successful assumption in the context of heavy-ion collisions [45]. For the coordinates on ΣX\Sigma_{X} we can then choose the longitudinal spacetime rapidity η=arctanh(X3/X0)\eta_{\parallel}={\rm arctanh}\,(X^{3}/X^{0}), transverse radius r=X12+X22r=\sqrt{X_{1}^{2}+X_{2}^{2}}, and the azimuthal angle ϕ=arctan(X2/X1)\phi=\arctan(X^{2}/X^{1}). The resulting normal vector is

v^μ=(coshη00sinhη).\displaystyle\hat{v}^{\mu}=\begin{pmatrix}\cosh\eta_{\parallel}\\ 0\\ 0\\ \sinh\eta_{\parallel}\end{pmatrix}\,. (59)

The induced metric and its inverse are for w1=η,w2=r,w3=ϕw_{1}=\eta_{\parallel},\,w_{2}=r,\,w_{3}=\phi

Gij=(τf20001000r2),Gij=(τf20001000r2),\displaystyle G_{ij}=\begin{pmatrix}\tau_{f}^{2}&0&0\\ 0&1&0\\ 0&0&r^{2}\end{pmatrix}\,,\qquad G^{ij}=\begin{pmatrix}\tau_{f}^{-2}&0&0\\ 0&1&0\\ 0&0&r^{-2}\end{pmatrix}\,, (60)

and the projection operator onto ΣX\Sigma_{X} is

Δμν=(sinh2η00sinhηcoshη01000010sinhηcoshη00cosh2η).\displaystyle\Delta^{\mu\nu}=\begin{pmatrix}\sinh^{2}\eta_{\parallel}&0&0&\sinh\eta_{\parallel}\cosh\eta_{\parallel}\\ 0&1&0&0\\ 0&0&1&0\\ \sinh\eta_{\parallel}\cosh\eta_{\parallel}&0&0&\cosh^{2}\eta_{\parallel}\end{pmatrix}\,. (61)

With this we get

X=τf,𝕏=(0r0).\displaystyle X_{\parallel}=\tau_{f}\,,\qquad\mathbb{X}_{\perp}=\begin{pmatrix}0\\ -r\\ 0\end{pmatrix}\,. (62)

We additionally assume that the system has a maximal radial extent R¯\bar{R}. The resulting integration measure is

dΣX=τΣrθ(R¯r)dηdrdϕ.\displaystyle d\Sigma_{X}=\tau_{\Sigma}r\theta(\bar{R}-r)\,d\eta_{\parallel}\,dr\,d\phi\,. (63)

Motivated by blast-wave model fits to experimental data at small beam energy [46], we choose R¯=8\bar{R}=8 fm and τΣ=5fm/c\tau_{\Sigma}=5\,{\rm fm/c}.

For the parametrization of the momenta, we take into account that the system only possesses rotational invariance in transverse direction. Thus, we use energy p0p_{0}, momentum in beam direction p3p_{3}, transverse momentum pT=p12+p22p_{T}=\sqrt{p_{1}^{2}+p_{2}^{2}} and azimuthal angle ϕp=arctan(p2/p1)\phi_{p}={\rm arctan}(p^{2}/p^{1}), which lead to

p=p0coshηp3sinhη,𝕡=(p0τfsinhηp3τfcoshηpTcos(ϕpϕ)pTrsin(ϕpϕ)).\displaystyle\begin{split}p_{\parallel}&=p_{0}\cosh\eta_{\parallel}-p_{3}\sinh\eta_{\parallel}\,,\\ \mathbb{p}_{\perp}&=\begin{pmatrix}p_{0}\tau_{f}\sinh\eta_{\parallel}-p_{3}\tau_{f}\cosh\eta_{\parallel}\\ -p_{T}\cos(\phi_{p}-\phi)\\ -p_{T}r\sin(\phi_{p}-\phi)\end{pmatrix}\,.\end{split} (64)

IV.4 Numerical results

Two-particle spectra are in general six-dimensional functions of the spatial momenta of the two particles. We express these in terms of the coordinates of the ambient Minkowski space. Since this is customary in experiments, we choose the center-of-momentum frame of the particle pair in beam direction, P3=0P_{3}=0. By fixing the remaining components of the average momentum PP, we can define the correlation in out, side, and long direction. To this end, we choose ϕP=0\phi_{P}=0, so that the average momentum points in 1-direction, resulting in

=(PT00).\displaystyle\mathbb{P}=\begin{pmatrix}P_{T}\\ 0\\ 0\end{pmatrix}\,. (65)

We then define the relative momenta

Δout=(ΔPT00),Δside=(0ΔPT0),Δlong=(00ΔP3).\displaystyle\mathbb{\Delta P}_{\rm out}=\begin{pmatrix}\Delta P_{T}\\ 0\\ 0\end{pmatrix}\,,\;\mathbb{\Delta P}_{\rm side}=\begin{pmatrix}0\\ \Delta P_{T}\\ 0\end{pmatrix}\,,\;\mathbb{\Delta P}_{\rm long}=\begin{pmatrix}0\\ 0\\ \Delta P_{3}\end{pmatrix}\,. (66)

In the following, we express the spectra in terms of these momenta, e.g., n1(𝕡,𝕢)=n1(+12Δ,12Δ)n1(,Δ)n_{1}(\mathbb{p}_{\perp},\mathbb{q}_{\perp})=n_{1}\bigg{(}\mathbb{P}_{\perp}+\frac{1}{2}\mathbb{\Delta P}_{\perp},\,\mathbb{P}_{\perp}-\frac{1}{2}\mathbb{\Delta P}_{\perp}\bigg{)}\equiv n_{1}(\mathbb{P},\mathbb{\Delta P}). Since our parametrization is in terms of the momenta in ambient-space coordinates, we also express the correlations in terms of these in the following.

With a specific parametrization of the momenta, we can address the validity of assuming that the emission function discussed in Sec. IV.1 only depends on the average momentum. From Eqs. (65) and (66) follows that the particle momenta ±Δ\mathbb{P}\pm\mathbb{\Delta P} have equal magnitude in side- and long-direction. This leads to vanishing relative on-shell energies in ambient-space coordinates, ΔP¯0\overline{\Delta P}_{0}. However, the relative energy on the hypersurface, ΔP¯=ΔP¯0coshηΔP3sinhη\overline{\Delta P}_{\parallel}=\overline{\Delta P}_{0}\cosh\eta_{\parallel}-\Delta P_{3}\sinh\eta_{\parallel}, is relevant for the correlations. This only leads to a non-vanishing contribution if both ΔP¯0\overline{\Delta P}_{0} and ΔP3\Delta P_{3} vanish. In our case, this is true for the side-correlation. Hence, only the side-correlation can be described by an emission function S(x,)S(x,\mathbb{P}_{\perp}) in Eq. (50), the emission functions in out and long directions also depend on the correction ΔS(x,,Δ)\Delta S(x,\mathbb{P}_{\perp},\mathbb{\Delta P}_{\perp}) in Eq. (51).

Refer to caption
Figure 5: Single-particle spectrum from Eq. (55). The parameters are the same as in Fig. 4.

To fix the remaining free parameters, we choose m=140m=140 MeV for the effective mass since it has been predicted in Ref. [5] that pions are sensitive to the moat regime, and are also commonly used to perform experimental HBT studies. We assume that the hypersurface ΣX\Sigma_{X} is defined by a fixed temperature of T=115T=115 MeV, which is consistent with the anticipated location of a moat regime in the QCD phase diagram [5, 20]. Unless otherwise stated, λ=202MeV\lambda=202\,{\rm MeV} and M=115MeVM=115\,{\rm MeV} for the parameters of the dispersion relation in Eq. (58). Since λ2M2>1\frac{\lambda^{2}}{M^{2}}>1 in this case, the system is in a moat regime for these parameters. This choice leads to |𝕡min|=100MeV|\mathbb{p}_{\rm min}|=100\,{\rm MeV} and Emin=110MeVE_{\rm min}=110\,{\rm MeV}.

Refer to caption
Refer to caption
Refer to caption
Figure 6: Two-particle spectrum n2(,Δ)n_{2}(\mathbb{P},\mathbb{\Delta P}) from Eq. (56) in out, side, and long direction in the normal phase. We denote P=||P=|\mathbb{P}| and ΔP=|Δ|\Delta P=|\mathbb{\Delta P}|.
Refer to caption
Refer to caption
Refer to caption
Figure 7: Two-particle spectrum n2(,Δ)n_{2}(\mathbb{P},\mathbb{\Delta P}) from Eq. (56) in out, side, and long direction in the moat regime. We denote P=||P=|\mathbb{P}| and ΔP=|Δ|\Delta P=|\mathbb{\Delta P}|.

In the following, we will always compare results in the normal phase to the moat regime. This is distinguished by the underlying dispersion relation,

normal phase:ω𝕡=𝕡2+m2moat regime:ω𝕡=Z(𝕡2)𝕡2+m2\displaystyle\begin{split}{\rm\emph{normal phase}}:&\quad\omega_{\mathbb{p}_{\perp}}=\sqrt{\mathbb{p}_{\perp}^{2}+m^{2}}\\ {\rm\emph{moat regime}}:&\quad\omega_{\mathbb{p}_{\perp}}=\sqrt{Z(\mathbb{p}_{\perp}^{2})\,\mathbb{p}_{\perp}^{2}+m^{2}}\end{split} (67)

with Z(𝕡2)Z(\mathbb{p}_{\perp}^{2}) specified at the beginning of Sec. IV.3 and the parameters given above.

IV.4.1 Single-particle spectrum

The single-particle spectrum n1(𝕡)n_{1}(\mathbb{p}) on ΣX\Sigma_{X} as a function of transverse momentum is shown in Fig. 5 at p3=0p_{3}=0 and ϕp=0\phi_{p}=0. The grey lines are for a particle in the normal phase and the yellow lines are in a moat regime. As also shown in Ref. [19], the spectrum of a particle in a moat regime shows a characteristic non-monotonic pTp_{T}-dependence. The initial increase of the spectrum with increasing momentum in the moat regime results from the decrease in energy in a moat regime. The location of the resulting peak is related to the wave number of the underlying spatial modulation, determined by the momentum at the bottom of the moat, 𝕡min\mathbb{p}_{\rm min}.

Refer to caption
Refer to caption
Refer to caption
Figure 8: HBT radii RoutR_{\rm out}, RsideR_{\rm side}, and RlongR_{\rm long} as functions of the average momentum in the normal phase (gray) and the moat regime (yellow).

IV.4.2 Two-particle spectrum

We have explicitly checked that the particle-antiparticle interference term is negligible against the particle-particle interference contribution, |n¯1(,Δ)|2|n1(,Δ)|2102\frac{|\bar{n}_{1}(\mathbb{P},\mathbb{\Delta P})|^{2}}{|n_{1}(\mathbb{P},\mathbb{\Delta P})|^{2}}\lesssim 10^{-2}. Hence, the last term in the two-particle spectrum n2(,Δ)n_{2}(\mathbb{P},\mathbb{\Delta P}), defined in Eq. (48), can be neglected.

In Fig. 6 we show the two-particle spectrum within our qualitative approximations in out, side, and long direction in the normal phase. This has to be compared to Fig. 7, where we show these correlations in the moat regime. In the normal phase, the two-particle spectra decrease monotonically with increasing average pair momentum |P||P|. In the moat regime n2n_{2} is a non-monotonic function of |P||P|, with a pronounced peak at nonzero |P||P| and small relative pair momentum |ΔP||\Delta P|, which continues as a ridge with increasing |ΔP||\Delta P|. Similar to the single-particle spectrum in Fig. 5, the location of this peak is related to the wavenumber of the underlying spatial modulation. The reason is that the average momentum is put on-shell in the particle-particle interference term Eq. (56), so the production of particle pairs with relative momenta close to the minimum of the energy ω\omega_{\mathbb{P}_{\perp}}, i.e., |||min||\mathbb{P}|\approx|\mathbb{P}_{\rm min}|, is enhanced. Consequently, also the correlation is enhanced. For the same reason, correlations in the normal phase peak at zero |||\mathbb{P}|. The moat regime shifts this peak to nonzero momentum. This is induced by the in-medium modifications of the spectral function of the particles, cf. Eq. (45).

Our result is qualitatively similar to the result in Ref. [19], where such a peak also arises in the two-particle correlation generated through thermodynamic fluctuations. In fact, the underlying reason for the non-monotonic structure of correlations is exactly the same here as in Ref. [19]. We also note that the peaks seen in the two-particle correlations are manifestly different from the structures that arise from final-state Coulomb interactions [47], since these occur in the relative momentum dependence, while our signal is in the average momentum dependence.

IV.4.3 HBT radii

To extract the HBT radii, the normalized two-particle correlation C(,Δ)C(\mathbb{P},\mathbb{\Delta P}), defined in Eq. (2), is considered. At vanishing relative momentum C(,𝟘)=2C(\mathbb{P},\mathbb{0})=2. This is known as the Goldhaber-Goldhaber-Lee-Pais effect [23], and follows trivially from the definition of the interference term, Eq. (33). Consequently, the normalized correlations do not show peaks as a function of |||\mathbb{P}| at |Δ|=0|\mathbb{\Delta P}|=0, unlike n2n_{2}, see Fig. 7. However, ratios of the normalized correlations turn out to be very sensitive to the moat regime. They are shown in Fig. 2 in the normal phase and in Fig. 3 in the moat regime, and are discussed in Sec. II.

As as discussed in Sec. IV.1, since the emission function connects the spacetime structure of the source to the relative momentum dependence of the correlation, cf. Eq. (52), the inverse range of the correlations in Δ\mathbb{\Delta P} reflects the size of the region of homogeneity in 𝕩\mathbb{x}. The HBT radii extracted from the range of C(,Δ)C(\mathbb{P},\mathbb{\Delta P}) are interpreted as measuring the spatial extent of the region of homogeneity. We determine the HBT radii through

R(||)=1|Δ(||)|,\displaystyle R(|\mathbb{P}|)=\frac{1}{\big{|}\mathbb{\Delta P}^{*}(|\mathbb{P}|)\big{|}}\,, (68)

where Δ\mathbb{\Delta P}^{*} is the relative momentum where the correlation falls off to half its maximal size, i.e.,

C(,Δ)=3/2,\displaystyle C\big{(}\mathbb{P},\mathbb{\Delta P}^{*}\big{)}=3/2\,, (69)

where we used that C(,𝟘)=2C(\mathbb{P},\mathbb{0})=2. In Fig. 8 we compare the radii in out, side, and long direction in the moat regime to the normal phase. We see that while the radii in a moat regime are clearly different from the ones in the normal phase, the details depend strongly on the direction of the relative momentum. RoutR_{\rm out} is a bit increased in the moat regime, but most strikingly shows a pronounced dip around |||min|=100|\mathbb{P}|\approx|\mathbb{P}_{\rm min}|=100 MeV. A similar dipping behavior is observed in RsideR_{\rm side}, but it is far less pronounced. In this case, the radius beyond the shallow dip is similar to the one in the normal phase. Instead of a dip, RlongR_{\rm long} shows a peak in this region. Thus, the long correlation falls of more rapidly in a moat regime if the average momenta of the particles are close to the minimal one. In summary, also HBT radii are sensitive to the moat regime.

While HBT radii can always be computed in principle, whether they have the desired physical meaning in the present case is questionable. First, we have shown in Sec. IV.1 that particle-particle interference cannot in general be described by an emission function that allows for an interpretation as phase-space distribution of particles. Second, our correlation functions show a rather strong dependence on the average momentum. Hence, the smoothness approximation which underlies the physical interpretation of HBT radii [48] might not be valid here.

V Conclusion

We have shown that HBT correlations are sensitive probes of a moat regime in heavy-ion collisions. To facilitate this, we developed a field-theoretical formalism that relates nn-particle spectra to 2n2n-point functions of quantum fields on curved hypersurfaces of spacetime. This way, in-medium modifications of particles on hypersurfaces of the expanding medium created in a heavy-ion collision can leave signatures in the spectra of these particles. We have shown that in the Gaussian approximation, these signatures arise in two-particle correlations through the spectral functions of the particles in the medium.

Our qualitative study shows that correlation functions of two identical bosons in a moat regime show characteristic peaks at nonzero average pair momentum. This is in contrast to a normal phase without spatial modulations, where these correlations always peak at zero average momentum. This effect can be interpreted as a result of enhanced particle production: Since particles have less energy at the bottom of the moat, they are more abundantly produced carrying the momentum that minimizes their energy, thereby directly enhancing also correlations at this momentum. This has already been pointed out in Ref. [19], where thermodynamic correlations have been analyzed, but a general formalism to describe particle spectra was still lacking.

We have shown that HBT radii, which are extracted from normalized two-particle correlations, are modified in a moat regime as well. It turned out that ratios of these normalized two-particle correlations are particularly sensitive probes of a moat regime. Due to their distinguished role in interferometry, normalized correlations of identical bosons, such as pions, are extensively studied in heavy-ion collisions. Their ratios have not been studied yet, but are promising observables to detect a moat regime.

We conclude that momentum-dependent particle correlations can be used as experimental probes of novel phases in QCD in heavy-ion collisions. Based on first indications of a moat regime in the QCD phase diagram found in Ref. [5], we expect these signatures to appear in lower-energy collisions, which could be within reach, e.g., of the STAR Fixed-Target program or HADES, and future experiments such as CBM at FAIR.

Acknowledgements.
We are grateful to Eduardo Grossi and Mike Lisa for valuable discussions. Upon completion of this work, we learned that Kenji Fukushima and Yoshimasa Hidaka have done an independent study of HBT in a moat regime. This work is supported by the Deutsche Forschungsgemeinschaft (DFG, German Research Foundation) through the Collaborative Research Center TransRegio CRC-TR 211 “Strong- interaction matter under extreme conditions” – project number 315477589 – TRR 211. The work of D.H.R. is supported by the State of Hesse within the Research Cluster ELEMENTS (Project ID 500/10.006).

Appendix A Particle-antiparticle interference

Here we derive the expression for particle-antiparticle interference, which is the third contribution in Eq. (48). We have with the relative and average momenta defined in Eq. (36):

n¯1(𝕡,𝕢)=12limxy𝑑Σμ𝑑Σνeip¯xeiq¯y(μx+ip¯μ)(νy+iq¯ν)×ϕ(x)ϕ(y)=12limΔX0𝑑ΣX𝑑ΣΔXe2iP¯Xei2ΔP¯ΔX×(14(X)2(ΔX)2+iΔP¯XiΔP¯+14ΔP¯2P¯2)×ϕ(X+12ΔX)ϕ(X12ΔX).\displaystyle\begin{split}&\bar{n}_{1}(\mathbb{p}_{\perp},\mathbb{q}_{\perp})\\ &\;=-\frac{1}{2}\lim_{x_{\parallel}\rightarrow y_{\parallel}}\int\!d\Sigma^{\mu}\,d\Sigma^{\prime\nu}\,e^{-i\bar{p}\cdot x}e^{-i\bar{q}\cdot y}\,(\partial_{\mu}^{x}+i\bar{p}_{\mu})(\partial_{\nu}^{y}+i\bar{q}_{\nu})\\ &\;\quad\times\langle\phi(x)\phi(y)\rangle\\ &\;=-\frac{1}{2}\lim_{\Delta X_{\parallel}\rightarrow 0}\int\!d\Sigma_{X}\,d\Sigma_{\Delta X}\,e^{-2i\overline{P}\cdot X}e^{-\frac{i}{2}\overline{\Delta P}\cdot\Delta X}\\ &\;\quad\times\bigg{(}\frac{1}{4}\big{(}\partial_{\parallel}^{X}\big{)}^{2}-\big{(}\partial^{\Delta X}_{\parallel}\big{)}^{2}+i\overline{\Delta P}_{\parallel}\partial_{\parallel}^{X}-i\overline{\Delta P}_{\parallel}+\frac{1}{4}\overline{\Delta P}_{\parallel}^{2}-\overline{P}_{\parallel}^{2}\bigg{)}\\ &\;\quad\times\bigg{\langle}\phi\bigg{(}X+\frac{1}{2}\Delta X\bigg{)}\phi\bigg{(}X-\frac{1}{2}\Delta X\bigg{)}\bigg{\rangle}\,.\end{split} (70)

Carrying out the Wigner transformation yields

n¯1(𝕡,𝕢)=12dΣXe2iP¯XdΔP2π[14(X)2+iP¯X+14(ΔP+ΔP¯)2P~2]×[F(X,12ΔP)12ρ(X,12ΔP)].\displaystyle\begin{split}&\bar{n}_{1}(\mathbb{p}_{\perp},\mathbb{q}_{\perp})\\ &\;=-\frac{1}{2}\int\!d\Sigma_{X}\,e^{-2i\overline{P}\cdot X}\int\!\frac{d\Delta P_{\parallel}}{2\pi}\,\bigg{[}\frac{1}{4}\big{(}\partial_{\parallel}^{X}\big{)}^{2}+i\overline{P}_{\parallel}\partial_{\parallel}^{X}\\ &\;\quad+\frac{1}{4}\big{(}\Delta P_{\parallel}+\overline{\Delta P}_{\parallel}\big{)}^{2}-\widetilde{P}_{\parallel}^{2}\bigg{]}\\ &\;\quad\times\bigg{[}F\bigg{(}X,\frac{1}{2}\Delta P\bigg{)}-\frac{1}{2}\rho\bigg{(}X,\frac{1}{2}\Delta P\bigg{)}\bigg{]}\,.\end{split} (71)

For this contribution the relative momentum is associated with the relative location and hence enters the Wigner-transformed two-point functions.

Assuming local thermodynamic equilibrium and that gradients in the average location are negligible, we find in analogy to Eq. (45):

n¯1(𝕡,𝕢)=18dΣXe2iP¯XdΔP2π[(ΔP+ΔP¯)24P¯2]f(X,12ΔP)ρ(X,12ΔP).\displaystyle\begin{split}&\bar{n}_{1}(\mathbb{p}_{\perp},\mathbb{q}_{\perp})\\ &\quad=-\frac{1}{8}\int\!d\Sigma_{X}\,e^{-2i\overline{P}\cdot X}\int\!\frac{d\Delta P_{\parallel}}{2\pi}\,\Big{[}\big{(}\Delta P_{\parallel}+\overline{\Delta P}_{\parallel}\big{)}^{2}\\ &\quad\quad-4\overline{P}_{\parallel}^{2}\Big{]}\;f\bigg{(}X,\frac{1}{2}\Delta P\bigg{)}\,\rho\bigg{(}X,\frac{1}{2}\Delta P\bigg{)}\,.\end{split} (72)

References