This is experimental HTML to improve accessibility. We invite you to report rendering errors. Use Alt+Y to toggle on accessible reporting links and Alt+Shift+Y to toggle off. Learn more about this project and help improve conversions.
HTML conversions sometimes display errors due to content that did not convert correctly from the source. This paper uses the following packages that are not yet supported by the HTML conversion tool. Feedback on these issues are not necessary; they are known and are being worked on.
failed: orcidlink
failed: multibib
Authors: achieve the best HTML results from your LaTeX submissions by following these best practices.
Can the strong interactions between hadrons be determined using femtoscopy?
Evgeny Epelbaum\orcidlink0000-0002-7613-0210
evgeny.epelbaum@ruhr-uni-bochum.deRuhr-Universität Bochum, Fakultät für Physik und Astronomie, Institut für Theoretische Physik II, D-44780 Bochum, Germany
Sven Heihoff
sven.heihoff@ruhr-uni-bochum.deRuhr-Universität Bochum, Fakultät für Physik und Astronomie, Institut für Theoretische Physik II, D-44780 Bochum, Germany
Ulf-G. Meißner\orcidlink0000-0003-1254-442X
meissner@hiskp.uni-bonn.deHelmholtz-Institut für Strahlen- und Kernphysik,
Rheinische Friedrich-Wilhelms Universität Bonn, D-53115 Bonn, Germany
Bethe Center for Theoretical Physics, Rheinische Friedrich-Wilhelms Universität Bonn, D-53115 Bonn, Germany
Center for Science and Thought, Rheinische Friedrich-Wilhelms Universitäat Bonn, D-53115 Bonn, Germany
Institute for Advanced Simulation (IAS-4), Forschungszentrum Jülich, D-52425 Jülich, Germany
Peng Huanwu Collaborative Center for Research and Education,
International Institute for Interdisciplinary and Frontiers, Beihang University, Beijing 100191, China
Alexander Tscherwon
alexander.tscherwon@ruhr-uni-bochum.deRuhr-Universität Bochum, Fakultät für Physik und Astronomie, Institut für Theoretische Physik II, D-44780 Bochum, Germany
Abstract
In the last decades, femtoscopic measurements from heavy-ion collisions have become a popular tool to
investigate the strong interactions between hadrons. The key observables measured in such experiments
are the two-hadron momentum correlations, which depend on the production mechanism of hadron pairs
and the final-state interactions. Given the complexity of ultra-relativistic collision experiments,
the source term describing the production mechanism can only be modeled phenomenologically
based on numerous assumptions. The commonly employed approach for analyzing femtoscopic data
relies on the Koonin-Pratt formula, which relates the measured correlation functions with
the relative wave function of an outgoing hadron pair and a source term that
is assumed to be universal.
Here, we critically examine this universality assumption and show that for strongly interacting
particles such as nucleons,
the interpretation of femtoscopic measurements suffers from a potentially large intrinsic uncertainty.
We also comment on the ongoing efforts to explore three-body interactions using this experimental technique.
Introduction.—Femtoscopic measurements are becoming increasingly popular as a tool to study
hadronic interactions and constitute an integral part of the ongoing and upcoming experimental
programs at high-energy heavy-ion facilities such as RHIC, LHC, GSI/FAIR and J-PARC-HI. Originally
developed as an imaging technique to extract spatial and temporal characteristics of particle
production mechanisms in ultra-relativistic heavy-ion collisions [1, 2, 3, 4],
femtoscopy has nowadays evolved into a method for measuring the strong interactions between hadrons
[5, 6]. A traditional way of exploring hadron structure and dynamics
by means of scattering experiments aims at the determination of the corresponding on-shell scattering
amplitudes, which are unambiguously defined and experimentally measurable quantities. Unfortunately,
the short lifetime of hadronic resonances makes it difficult to perform scattering experiments beyond
the lightest hadrons. For example, there is a wealth of experimental data on nucleon-nucleon (NN)
scattering, while only a handful of/no data are available for hyperon-nucleon/hyperon-hyperon
scattering. Femtoscopic measurements from high-energy proton-proton (pp) or heavy-ion collisions
do not require preparing a beam of unstable particles and thus provide an attractive alternative
to scattering experiments. This, however, comes at the price of dealing with a complicated
initial state after hadronization, whose modeling becomes an integral part of the analysis
of the experimental data. After making several approximations and assumptions [4],
the two-particle correlation function in the center-of-mass system (cms)
is usually written in the form [1, 7, 4]
(1)
where is the relative momentum, is the source function and
is the relative wave function of the outgoing two-body state,
which coincides with the stationary solution of the scattering problem normalized to yield in the absence of the interaction, see e.g. Ref. [8].
This is the celebrated Koonin-Pratt formula, which is widely utilized to analyze femtoscopic measurements.
The pathway from experiment to interpretation can then be schematically summarized as follows
(see Refs. [5, 9] for details):
i.
Measurement of the correlation functions .
ii.
Modeling of the source function which is deemed to be universal.
Here, one usually assumes a spherically symmetric Gaussian form, whose size is extracted from
experimental data on using some model to describe the strong interaction. In
particular, is often fitted to pp correlation functions using the Reid Soft-Core [10]
or Argonne [11] potential models [12]. A more
sophisticated modeling of the source is described in Ref. [13].
iii.
Once the source function is fixed, Eq. (1)
allows one to probe hadronic interaction models using experimental data on .
It goes beyond the scope of this paper to address the validity of various assumptions and
approximations used to arrive at the formula (1), which are, in fact, well documented
in the literature [14, 15, 4, 8]. Rather, we would
like to point out that the way Eq. (1) is used to probe hadronic interactions as
outlined above suffers from a fundamental flaw: Combined with the universality assumption for
the source function , it implies the measurability of hadronic wave functions
and thus also of the corresponding interaction potentials. Yet, hadronic potentials are well known
not to represent observable quantities, see e.g. Ref. [16] for a recent
discussion. While interaction potentials can, at least in principle, be determined ab initio
using lattice QCD [17], their form depends on the choice of interpolating fields
for parametrizing hadrons in terms of valence quarks. This inherent scheme dependence does,
of course, not affect observable quantities like S-matrix, as can be shown by virtue of the
LSZ reduction formalism [18].
To make the essence of the problem more explicit while keeping the presentation simple, we
restrict ourselves to non-relativistic systems in the framework of quantum mechanics and
start with rewriting Eq. (1) in a more general form
(2)
The Koonin-Pratt formula (1) is the coordinate-space representation of
Eq. (2), subject to the restriction that the source term is local,
.
Eq. (2) makes it explicit that the correlation functions are invariant
under a change of basis in the Hilbert space induced by unitary transformations (UTs)
as required for any observable quantity:
(3)
However, it also shows that the source term cannot be assumed to be universal across interaction models,
and its form must be off-shell consistent with that of the interaction under consideration. Failure
to maintain off-shell consistency between the source and the wave function (or interaction
potential) by assuming leads to model dependence of the calculated correlation
functions. Given the quest for precision in femtoscopy experiments [6], it is important to
quantify model dependence in caused by the assumed universality of the source.
Gedankenexperiment.—For demonstration purposes, we focus here on two stable distinguishable
spinless particles interacting via a short-range force (but all arguments made here remain valid
for identical particles, nonvanishing spin and/or if the Coulomb interaction is included) and
perform a Gedankenexperiment by letting Alice and Bob analyze two-body correlations.
The interaction between particles considered by Alice coincides with the spin- projection
of the neutron-proton chiral effective field theory (EFT) potential at fifth order (N4LO+)
[19] with the cutoff parameter MeV. That is,
(4)
where and with and the initial and
final momenta in the cms, respectively, while is the orbital angular momentum. The corresponding
phase shifts in the partial waves are shown by black open circles in Fig. 1.
Note that we use natural units with throughout this paper.
Figure 1: (Color online). S-, P-, D- and F-wave phase shifts for the interaction
are shown with open circles as functions of the cms momentum. The overlapping blue dotted and
dashed lines show the phase shifts obtained from and , respectively.
For the source term, Alice adopts the commonly chosen Gaussian form
(5)
where with the momentum transfer. The radius
is set to fm.
The S-wave momentum-space matrix elements of the potential and the source term used by Alice and
specified in Eqs. (4) and (5) are visualized in the upper
row of Fig. 2.
Figure 2: (Color online). S-wave momentum-space matrix elements of the potential in units of
GeV-2 (left column) and of the dimensionless source term (right column) as chosen
by Alice (upper row) and seen by Bob (middle and bottom rows).
The correlation functions are usually calculated in coordinate space. Under the commonly made
assumption that particles interact only in the state, Eq. (1) can,
for a spherically symmetric source function, be reduced to (see, e.g., Ref. [20]):
(6)
with the spherical Bessel function and the S-wave scattering wave function
is normalized according to
(7)
Here, is the S-wave phase shift.
The wave-function can be obtained from the half-shell T-matrix,
see e.g. Ref. [21].
However, to avoid numerical integrations of strongly oscillating
functions, it is more convenient to directly compute the matrix element in Eq. (2)
in the partial-wave momentum-space basis. For a spherically symmetric
source, we find (see also Ref. [22]):
(8)
where we have introduced a short-hand notation ,
and
(9)
Here, is the reduced mass and denotes the Cauchy principal value integral.
The half-shell K-matrix can be easily calculated for arbitrary short-range interactions by
solving the Lippmann-Schwinger integral equation
(10)
using standard numerical methods. The on-shell K-matrix is related to the S-matrix via
.
We note in passing that Eq. (8) can be brought to the form of Eq. (6)
if the source term is local and the interaction vanishes in all channels.
The correlation function calculated by Alice using Eq. (8) is shown by open circles
in Fig. 3.
We have also checked these results by using Eq. (1) directly.
Figure 3: (Color online). Correlation functions calculated by Alice and Bob. The results of
Alice are shown by open circles. Red dash-dotted and dash-double-dotted lines show the
correlation functions calculated by Bob using the interactions and
, respectively, and the source term (assumed to be universal).
The overlapping blue dotted and dashed lines show the results of Bob using the
properly transformed source terms and , respectively.
On the other hand, Bob performs his analysis using a different basis in the Hilbert space with
, where the unitary operator is taken
as a rank-1 separable form
with the normalization constant fixed from .
The parameter specifies the inverse range of the transformation, while
gives the position of the node in the
profile function . The interaction in Bob’s convention has the form
(13)
where is the cms momentum operator. It can easily be calculated numerically
in momentum space. Notice that the considered transformation acts only in the S-wave.
To be specific, consider two different transformations corresponding to the parameters
, (set-I) and ,
(set-II).
These transformations significantly affect the S-wave momentum-space matrix elements of the
interaction as shown in Fig. 2. Yet, physical observables are, of course, independent of
the choice of basis in the Hilbert space. In particular, we have verified that the phase shifts
calculated from ,
and agree with each other to machine precision,
see Fig. 1.
In contrast, the correlation functions calculated by Bob under the assumption of the universal
source term, , are as expected
different from the result found by Alice, as depicted by the red lines in Fig. 3.
To restore the result for obtained by Alice, the source term
needs to be brought into Bob’s convention by applying the UT, , see the right column of Fig. 2 and blue lines in Fig. 3.
Scheme dependence in chiral EFT.—The above example shows that off-shell ambiguities of
the interaction potentials can result in a large model dependence of the correlation function
if the source term is assumed to be universal. On the other hand,
realistic models of hadronic interactions are constrained by physical principles like,
e.g., pion-exchange dominance at large distances and often share similarities when it comes to
modeling of the short-distance behavior. The remaining off-shell ambiguities may therefore be
expected to be less pronounced in practice.
In this context, it is instructive to examine how scheme dependence manifests itself in
chiral EFT for nuclear forces,
[24, 25, 26],
the most extensively studied and best understood hadronic interactions. Leaving aside the
(significant) ambiguity from the choice of the regulator, the inherent scheme dependence in
nuclear chiral EFT first shows up at fourth expansion order (N3LO), stemming from the
long-range relativistic corrections that depend on two arbitrary phases [27, 28]
and three short-range interactions contributing to the 1S0, 3S1 and 3S1-3D1 potentials
[19]. The corresponding five parameters can be chosen arbitrarily, subject to
the naturalness constraint, and their values can be changed by applying suitable UTs, see
Supplemental Material [29] for explicit expressions. Importantly, such UTs also induce
many-body interactions and exchange currents.
This shows, once again, that various scheme-dependent quantities such as two-body interactions,
three-body forces (3BFs) and exchange currents must be used consistently with each other
to avoid an uncontrollable model dependence.
The above power-counting-based arguments point towards a mild scheme-dependence of NN interactions
in chiral EFT. However, the situation is different for 3BFs, which provide small but important
corrections to the dominant pairwise forces and remain a challenging frontier in nuclear
physics [30], see also Refs. [31, 32, 33]
for recent attempts to explore 3BFs through femtoscopy. 3BFs first appear at third order (N2LO)
in chiral EFT, and thus are much more sensitive to the above mentioned off-shell ambiguities.
To illustrate this point we have generated a set of NN N4LO+ potentials corresponding
to different choices of the off-shell short-range interactions, see [29] for details.
These potentials are nearly phase-equivalent in the two-body sector and equally valid from the EFT
point of view. However, they lead to vastly different predictions for three-nucleon observables
as visualized in Fig. 4, which illustrates the well-known
inherent scheme dependence of
[34]
3BFs [35]. In particular, the required 3BF contributions to the 3H
binding energy can be both attractive and repulsive depending on the off-shell behavior
of the employed NN interaction.
Figure 4: (Color online). Neutron-deuteron total cross section (left) and the 3H binding
energy (right) calculated using the N4LO+ NN potential of Ref. [36]
(orange lines). Light-shaded blue bands show the results from the phase-equivalent but
off-shell differing N4LO+ NN potentials as explained in the text.
Discussion and conclusions.—Hadronic correlations measured in ultra-relativistic
collisions are sensitive to the strong interactions. However, probing final state interactions
by means of the Koonin-Pratt formula violates basic principles of quantum mechanics if the source
model is regarded to be universal. Using an example of two strongly interacting distinguishable
particles, we have shown that off-shell ambiguities in the interaction can then translate into
a significant model dependence for . It is important to emphasize that the problematic
universality assumption is essential as its relaxation sacrifices the predictive power of the
femtoscopy approach.
The sensitivity of to the off-shell behavior of the strong force
decreases for source radii being large compared to the
interaction range
since the large-distance behavior of the wave function is unambiguously determined by the S-matrix [37, 8, 4].
The interpretation of in
terms of the average
then becomes no more inherently problematic.
Large -values, however, also reduce the strength of femtoscopic signals. The source size
in our example is, in fact, comparable to or even larger than those typically used in the literature
[5, 6, 13, 9, 32].
Even smaller values for were obtained recently using a precise pion-kaon interaction [38].
Finally, we also discussed off-shell ambiguities of nuclear interactions in chiral EFT. While
scheme-dependent contributions to the NN force are suppressed by power counting, the 3BF is
highly sensitive to the off-shell components of the NN potentials. The lack of the off-shell
consistency in femtoscopy analyses, therefore, raises concerns about the feasibility of
extracting 3BFs from heavy-ion collisions as anticipated in
Refs. [31, 32, 33].
Acknowledgements.
One of the authors (EE) is grateful to the organizers of the “Three-body femtoscopy meeting” in Prague, March 4-6, 2025, for their hospitality and appreciates useful discussions with Raffaele Del Grande, Laura Fabbietti and Alejandro Kievsky. This work has been supported by the European Research Council (ERC) under the European Union’s Horizon 2020 research and innovation programme (grant agreement No. 885150 and No. 101018170), by the MKW NRW under the funding code NW21-024-A, by JST ERATO (Grant No. JPMJER2304), by JSPS KAKENHI (Grant No. JP20H05636) and by the Chinese Academy of Sciences (CAS) President’s International Fellowship Initiative (PIFI) (Grant No. 2025PD0022).
References
[1]
S. E. Koonin.
Proton Pictures of High-Energy Nuclear Collisions.
Phys. Lett. B, 70:43–47, 1977.
doi:10.1016/0370-2693(77)90340-9.
[2]
W. Bauer, C. K. Gelbke, and S. Pratt.
Hadronic interferometry in heavy ion collisions.
Ann. Rev. Nucl. Part. Sci., 42:77–100, 1992.
doi:10.1146/annurev.ns.42.120192.000453.
[3]
S. Pratt.
What we are learning from correlation measurements.
Nucl. Phys. A, 638:125–134, 1998.
doi:10.1016/S0375-9474(98)00407-2.
[5]
L. Fabbietti, V. Mantovani Sarti, and O. Vazquez Doce.
Study of the Strong Interaction Among Hadrons with Correlations at
the LHC.
Ann. Rev. Nucl. Part. Sci., 71:377–402, 2021.
arXiv:2012.09806,
doi:10.1146/annurev-nucl-102419-034438.
[6]
Alice Collaboration et al.
Unveiling the strong interaction among hadrons at the LHC.
Nature, 588:232–238, 2020.
[Erratum: Nature 590, E13 (2021)].
arXiv:2005.11495,
doi:10.1038/s41586-020-3001-6.
[7]
R. Lednicky and V. L. Lyuboshits.
Final State Interaction Effect on Pairing Correlations Between
Particles with Small Relative Momenta.
Yad. Fiz., 35:1316–1330, 1981.
[10]
Roderick V. Reid, Jr.
Local phenomenological nucleon-nucleon potentials.
Annals Phys., 50:411–448, 1968.
doi:10.1016/0003-4916(68)90126-7.
[11]
Robert B. Wiringa, V. G. J. Stoks, and R. Schiavilla.
An Accurate nucleon-nucleon potential with charge independence
breaking.
Phys. Rev. C, 51:38–51, 1995.
arXiv:nucl-th/9408016, doi:10.1103/PhysRevC.51.38.
[12]
D. L. Mihaylov, V. Mantovani Sarti, O. W. Arnold, L. Fabbietti, B. Hohlweger,
and A. M. Mathis.
A femtoscopic Correlation Analysis Tool using the Schrödinger
equation (CATS).
Eur. Phys. J. C, 78(5):394, 2018.
arXiv:1802.08481,
doi:10.1140/epjc/s10052-018-5859-0.
[14]
S. Pratt.
Validity of the smoothness assumption for calculating two-boson
correlations in high-energy collisions.
Phys. Rev. C, 56:1095–1098, 1997.
doi:10.1103/PhysRevC.56.1095.
[18]
H. Lehmann, K. Symanzik, and W. Zimmermann.
On the formulation of quantized field theories.
Nuovo Cim., 1:205–225, 1955.
doi:10.1007/BF02731765.
[19]
P. Reinert, H. Krebs, and E. Epelbaum.
Semilocal momentum-space regularized chiral two-nucleon potentials
up to fifth order.
Eur. Phys. J. A, 54(5):86, 2018.
arXiv:1711.08821,
doi:10.1140/epja/i2018-12516-4.
[21]
Johann Haidenbauer and Ulf-G. Meißner.
Exploring the +p interaction by measurements of
the correlation function.
Phys. Lett. B, 829:137074, 2022.
arXiv:2109.11794,
doi:10.1016/j.physletb.2022.137074.
[22]
M. Albaladejo, A. Feijoo, J. Nieves, E. Oset, and I. Vidaña.
Femtoscopy correlation functions and mass distributions from
production experiments.
Phys. Rev. D, 110(11):114052, 2024.
arXiv:2410.08880,
doi:10.1103/PhysRevD.110.114052.
[23]
P. U. Sauer.
Can the charge symmetry of nuclear forces be confirmed by
nucleon-nucleon scattering experiments?
Phys. Rev. Lett., 32:626–630, 1974.
doi:10.1103/PhysRevLett.32.626.
[24]
Evgeny Epelbaum, Hans-Werner Hammer, and Ulf-G. Meißner.
Modern Theory of Nuclear Forces.
Rev. Mod. Phys., 81:1773–1825, 2009.
arXiv:0811.1338,
doi:10.1103/RevModPhys.81.1773.
[26]
Evgeny Epelbaum, Hermann Krebs, and Patrick Reinert.
High-precision nuclear forces from chiral EFT: State-of-the-art,
challenges and outlook.
Front. in Phys., 8:98, 2020.
arXiv:1911.11875,
doi:10.3389/fphy.2020.00098.
[28]
V. Bernard, E. Epelbaum, H. Krebs, and U.-G. Meißner.
Subleading contributions to the chiral three-nucleon force II:
Short-range terms and relativistic corrections.
Phys. Rev. C, 84:054001, 2011.
arXiv:1108.3816,
doi:10.1103/PhysRevC.84.054001.
[29]
See Supplemental Material at [URL].
[30]
Shimpei Endo, Evgeny Epelbaum, Pascal Naidon, Yusuke Nishida, Kimiko Sekiguchi,
and Yoshiro Takahashi.
Three-body forces and Efimov physics in nuclei and atoms.
Eur. Phys. J. A, 61(1):9, 2025.
arXiv:2405.09807,
doi:10.1140/epja/s10050-024-01467-4.
[33]
A. Kievsky, E. Garrido, M. Viviani, L. E. Marcucci, L. Serksnyte, and
R. Del Grande.
nnn and ppp correlation functions.
Phys. Rev. C, 109(3):034006, 2024.
arXiv:2310.10428,
doi:10.1103/PhysRevC.109.034006.
[34]
W. P. Abfalterer et al.
Inadequacies of the nonrelativistic 3N Hamiltonian in describing the
n + d total cross section.
Phys. Rev. Lett., 81:57–60, 1998.
doi:10.1103/PhysRevLett.81.57.
[35]
W. N. Polyzou and W. Glöckle.
Three-body interactions and on-shell equivalent two-body
interactions.
Few Body Syst., 9(2-3):97–121, 1990.
doi:10.1007/BF01091701.
[36]
P. Reinert, H. Krebs, and E. Epelbaum.
Precision determination of pion-nucleon coupling constants using
effective field theory.
Phys. Rev. Lett., 126(9):092501, 2021.
arXiv:2006.15360,
doi:10.1103/PhysRevLett.126.092501.
[37]
M. Gmitro, J. Kvasil, R. Lednicky, and V. L. Lyuboshits.
On the Sensitivity of Nucleon-nucleon Correlations to the Form of
Short Range Potential.
Czech. J. Phys. B, 36:1281, 1986.
doi:10.1007/BF01598029.
[38]
Miguel Albaladejo, Alejandro Canoa, Juan Nieves, Jose Ramón Peláez, Enrique
Ruiz-Arriola, and Jacobo Ruiz de Elvira.
The role of chiral symmetry and the non-ordinary
nature in femtoscopic correlations.
3 2025.
arXiv:2503.19746.
I Supplemental Material
I.1 Scheme dependence in nuclear chiral EFT and phase equivalent NN potentials at N4LO+
In this Supplemental Material we focus on the off-shell ambiguities in the chiral nuclear interactions and describe the construction of the phase-equivalent but off-shell differing N4LO+ NN potentials.
As emphasized in the main text, scheme dependence is an inherent
feature of nuclear interactions. Already the starting point for the
derivation of the nuclear interactions, the effective chiral Lagrangian
truncated at a given order in the derivative expansion, features a
certain degree of ambiguity as reflected in the freedom to perform
field redefinitions and apply equations of motion to eliminate various
terms. The derivation of nuclear potentials requires
off-the-energy-shell extensions of the (few-nucleon-irreducible) parts
of the scattering amplitudes and introduces additional scheme
dependence. For a description of various methods to derive
interactions in chiral EFT see Ref. [26] and
references therein. A novel approach using the path integral
formalism, which is capable of dealing with regularized (non-local)
effective Lagrangians [2], has been proposed in
Ref. [3].
Here, we focus on the expressions for the NN potentials, 3NFs and current operators given in Refs. [4, 5, 6, 7, 8, 9, 10, 28, 11, 12, 13, 14, 15, 16, 17, 18], which were obtained
using the method of unitary transformations (UTs). In this approach,
one performes a UT of the effective pion-nucleon Hamiltonian to
decouple the states on the Fock space involving pions from the purely
nucleonic ones. This is achieved in a perturbative framework by
utilizing the standard chiral expansion [4, 9]. To arrive at renormalized
expressions for the 3BFs and current operators it was necessary,
starting from N3LO, to
systematically exploit the unitary ambiguity of the nuclear
interactions by considering a broad class of transformations on the
nucleonic subspace of the Fock space starting from N3LO
[8, 10, 11, 14, 15, 16, 17, 18]. The corresponding unitary phases were then chosen in
such a way that the resulting potentials remain finite after
renormalization (using dimensional regularization). For the considered
class of UTs, the resulting static expressions for the nuclear forces
at N3LO and N4LO turn out to be determined unambiguously, while
the leading relativistic corrections depend on two arbitrary real
parameters , corresponding to the UTs
[14]
(S1)
with the anti-Hermitian generators given by
(S2)
where and are the projection operators on the purely nucleonic states and states involving a single pion, respectively, is the pion energy, is the leading-order pion-nucleon vertex proportional to the nucleon axial-vector coupling constant , is the leading relativistic correction to that vertex with denoting the nucleon mass, while refers to the kinetic energy operator of the nucleons.
The UT in Eq. (S1) induces scheme-dependent contributions to the nuclear force of the form
(S3)
where the ellipses refer to terms beyond N4LO and denotes the leading contribution to the NN force
stemming from the one-pion exchange and derivative-less contact
interactions, . Evidently, the commutator with the kinetic
energy term generates contributions to the
one-pion exchange potential, while the one with
induces the corrections to the NN and three-nucleon two-pion
exchange potentials [28]. The
resulting unitary ambiguities in the NN potential are well known and
discussed in detail in Ref. [27]. Notice further that
when considering reactions involving external electroweak probes, the
unitary transformation in Eq. (S1) must be applied to the
corresponding single-nucleon charge and current operators, yielding
scheme-dependent two-nucleon contributions
[19, 14, 15]. Approximate
independence of electromagnetic NN observables on the unitary phases
requires using consistent expressions for nuclear
potentials and current operators that correspond to the same values of
and has been explicitly verified in Refs. [20, 21].
In addition to the long-range UTs in Eq. (S1), nuclear
potentials at N3LO feature the short-range ambiguity related to the UTs
(S4)
where are arbitrary real parameters and the short-range
anti-Hermitian NN operators are given by
[19]
(S5)
Here, and denote the incoming and outgoing momenta of the nucleon while
is the breakdown scale of chiral EFT, respectively.
Note that in line with the commonly used notation, we
do not show in Eq. (I.1) the overall factor of
and the -function corresponding to the total
momentum conservation. We further emphasize that we do not consider
here short-ranged UTs, whose generators vanish in the NN cms
[22].
In a full analogy to Eq. (S3),
the induced contributions to the nuclear forces have the form
(S6)
In the two-nucleon system, the commutator with leads to short-range NN operators that can be absorbed into a redefinition of the contact interactions. On the other hand, the commutator with generates purely off-shell short-range interactions in the NN 1S0, 3S1 and 3S1- 3D1 partial waves. Up to N4LO, the contact interactions in these channels have the form
(S7)
where are the LO low-energy constants (LECs), are the next-to-leading order (NLO) LECs while
and are the N3LO LECs. Notice that the
contact interactions proportional to ,
and vanish in
the on-shell kinematics with and thus cannot be
determined reliably from fits to NN scattering data. This can also be
understood from a different perspective by observing that the
considered short-ranged UTs generate the NN contact interactions whose
form is identical to those proportional to when the
commutator in Eq. (S6) is evaluated with . This implies that the contact terms proportional to the
LECs can be rotated away by means of a suitably
chosen UT. Accordingly, the NN potentials of
Refs. [19, 36] were constructed using
the scheme with
(S8)
It is important to emphasize that the considered UTs also induce 3BFs
when evaluating the commutator in Eq. (S6) with [19, 23]. This
demonstrates, once again, that 3BFs can only be meaningfully defined
in conjunction with the NN potential, i.e., within a given scheme
fixed by a specific off-shell behavior of the NN interaction.
The choice specified in Eq. (S8) and adopted in
Refs. [19, 36] is obviously just one
possibility out of infinitely many. From the EFT point of view, any values of are acceptable provided the LECs of the NN contact interactions are of natural size.
To illustrate scheme dependence of the 3BF in chiral EFT, we have
generated a set of semi-local momentum-space regularized (SMS) NN
potentials [19, 36, 24],
which are nearly phase equivalent but differ by the choices of the
off-shell LECs , and . Specifically, we consider the
values
(S9)
in units of GeV-6 used in Ref. [19] and employ the
cutoff value of MeV. These units coincide with the
natural units of used in
Refs. [25, 26],
where and denote the pion decay
constant and the breakdown scale of chiral EFT, respectively, if one
sets MeV.
To ensure approximate phase equivalence of the resulting potentials,
we stay at the highest available EFT order N4LO+, which provides
a sufficient flexibility to achieve a statistically perfect
description of NN data up through the pion production threshold, see
Refs. [19, 24] for details, and
restrict ourselves to the cutoff value of MeV.
The assumed range of values for and
can be considered as fairly
conservative with regard to the naturalness assumption. For example,
with the off-shell LECs switched off, the resulting N3LO LECs
fulfill GeV-6 for MeV
[19]. For the LEC , the
values turn out to cause
unrealistically strong modifications of the deuteron D-state
probability, which would render the resulting potentials impractical
for many-body applications [26]. We have therefore
limited ourselves to the values .
For each of the 26 combinations of the off-shell LECs in
Eq. (S9) beyond the already available one
corresponding to Eq. (S8), we have fitted the remaining LECs
accompanying the NN contact interactions to the neutron-proton and
proton-proton scattering data up to MeV, following
the same protocol and using the same database as employed in
Refs. [36, 27].
We emphasize that when changing the values of the off-shell LECs
and/or the LECs beyond the considered expansion order, the LECs
accompanying contact interactions at all orders must be refitted. This
procedure corresponds to implicit renormalization within
the considered EFT framework and
ensures that all results are given in terms of physical
parameters (i.e., experimental data for observables used in the fit)
rather than bare LECs, see Ref. [26] for details.
For all considered cases, we found an essentially perfect description of the
neutron-proton (np) and proton-proton scattering data as reflected in the
-values in the range of . These results provide an important consistency
check of our calculations and show, in particular, that the
contributions in Eq. (I.1) that have been neglected at the
considered accuracy level based on the power-counting arguments are
indeed numerically negligibly small.
Figure S1: (Color online). The results for the neutron-proton phase
shifts and mixing angles calculated using the 27 N4LO+
potentials with different choices of the off-shell LECs
specified in Eq. (S9) are shown by
light-shaded blue bands. Phase shifts and mixing angles obtained
using the potential from Ref. [36] with are shown by
orange lines. In all cases, the cutoff is set to MeV.
In Fig. S1, we show the resulting np phase shifts and
mixing angles in low partial waves. These results demonstrate that the
27 potentials can indeed be regarded as essentially phase
equivalent. The visible (but tiny) dependence of the
mixing angle on the off-shell LECs reflects the fact that
this particular observable is less well constrained by the
experimental data as compared to other phase shifts. For an
estimation of the EFT truncation uncertainty and other types of uncertainties see
Refs. [26, 24, 27].
The predictions for the
deuteron properties for the 27 N4LO+ potentials are collected in Table S1.
⋆The deuteron binding energy has been taken as input in the fit.
Table S1: Deuteron binding energy , asymptotic S-state
normalization , asymptotic D/S-state ratio , matter
radius , leading contribution to the quadrupole moment
and D-state probability obtained using the SMS N4LO+
potential of Ref. [36] and the 27 N4LO+
potentials with the off-shell LECs set according to Eq. (S9) are given in
the second and third columns, respectively. The cutoff is set to MeV.
The resulting values for the S-state normalization observable and
the asymptotic D/S-state ratio show, similarly to the phase
shifts, a very small sensitivity to the off-shell LECs and agree with
the experimental data within errors, see Ref. [24]
for the error analysis. On the other hand, the
matter radius (i.e., the expectation value of the relative
distance between the nucleons), the quadrupole moment
of the deuteron wave function and the D-state probability
are well known not to correspond to observable quantities, see
e.g. [31]. Not surprisingly, our
predictions for these quantities demonstrate a significant degree of scheme
dependence.
In contrast to and , the deuteron charge radius and quadrupole
moment related to the electric and quadrupole form factors
and , respectively, via and with denoting the deuteron mass are, of
course, observables. The scheme-dependent quantities and provide the bulk
contributions to the charge radius and the quadrupole moment as they
originate from the dominant single-nucleon charge density, but one
also needs to take into account two-body contributions to the charge
density operator, which are often referred to as “meson-exchange currents”.
These (scheme-dependent) meson-exchange currents must be chosen consistently
with the nuclear interactions to ensure that the resulting form
factors are unambiguous. Indeed, it is easy to
see that the unitary transformation in Eq. (S4) generates
short-range meson-exchange currents when acting on the single-nucleon
charge density [21], rendering the calculated form factors independent of the
arbitrary phases . Note that the spread in the obtained results for
and agrees qualitatively with the expected size of
meson-exchange contributions from the power-counting, see
Refs. [21] for details.
Finally, in Fig. S2, we show the results for the
np and neutron-deuteron (nd) total cross section in comparison
with the experimental data. As already evident from the results for
phase shifts, the np cross section comes out nearly identical for all
27 potentials with the maximal relative
differences below two-permille level. On the other hand, the nd cross section calculated using
the NN forces only shows a significant scheme dependence, as already visible from Fig. 4.
Figure S2: (Color online). Neutron-proton and
neutron-deuteron total cross section as a function of the cms
momentum . The upper and lower inset plots show
the relative cross section differences for nd
and np
scattering, respectively, defined as . Light-shaded blue bands (not visible for
neutron-proton scattering) are obtained from the 27
nearly phase-equivalent N4LO+ NN potentials and
illustrate inherent scheme dependence of nuclear interactions in chiral
EFT. Experimental data for the nd and np total cross section
are taken from Refs. [34] and [34], respectively.
The difference between the nd experimental data and calculations
utilizing a particular NN potential gives the contribution of the 3NF
needed to reproduce the cross section data. The large spread in the
predictions for three-nucleon observables from different but nearly
phase equivalent N4LO+ NN potentials, reflected in the
light-shaded blue bands in Figs. 4 and S2,
demonstrates the strong inherent scheme dependence of
the 3BF in the framework of chiral EFT.
The results for the nd total cross section and the 3H
binding energy shown in Figs. 4 and S2 agree well
with estimations based on the power counting, see
Ref. [32] for a related discussion. For example,
given the expansion parameter in chiral EFT of the order of in chiral EFT and the expectation value of the NN interaction for
3H of MeV, one may
expect the 3BF contribution to the 3H binding energy roughly of
the order of MeV, where we took into
account the first appearance of the 3BF at N2LO ().
Furthermore, the approximately constant deviations between the experimental data for
the nd total cross sections and the blue band comprising the results based on the
27 N4LO+ NN potentials in Fig. 4 translate into the relative
deviation that grows with energy and reaches about % at MeV, which corresponds to MeV. This pattern and the amount of underprediction of
are also well in line with the expected size of
the 3BF contributions based on the chiral power counting
[32]. In particular, the growing contributions of the
3NF at increasing energies, also seen in the context of the nuclear
equation of state, are consistent with the assumed EFT
expansion parameter .
Last but not least, we emphasize that the leading 3BF at N2LO was
found to increase the nd total cross section in Ref. [33], thereby
improving the agreement between theory and experiment.
References
[1]
[2]
H. Krebs and E. Epelbaum,
Toward consistent nuclear interactions from chiral Lagrangians. II. Symmetry preserving regularization,
Phys. Rev. C 110, no.4, 044004 (2024).
arXiv:2312.13932 [nucl-th],
doi:10.1103/PhysRevC.110.044004.
[5]
E. Epelbaum, W. Glöckle and U.-G. Meißner,
Nuclear forces from chiral Lagrangians using the method of unitary transformation. 2. The two nucleon system,
Nucl. Phys. A 671, 295-331 (2000).
arXiv:nucl-th/9910064 [nucl-th],
doi:10.1016/S0375-9474(99)00821-0.
[10]
V. Bernard, E. Epelbaum, H. Krebs and U.-G. Meißner,
Subleading contributions to the chiral three-nucleon force. I. Long-range terms,
Phys. Rev. C 77, 064004 (2008).
arXiv:0712.1967 [nucl-th],
doi:10.1103/PhysRevC.77.064004.
[13]
S. Kölling, E. Epelbaum, H. Krebs and U.-G. Meißner,
Two-pion exchange electromagnetic current in chiral effective field theory using the method of unitary transformation,
Phys. Rev. C 80, 045502 (2009).
arXiv:0907.3437 [nucl-th],
doi:10.1103/PhysRevC.80.045502.
[14]
S. Kölling, E. Epelbaum, H. Krebs and U.-G. Meißner,
Two-nucleon electromagnetic current in chiral effective field theory: One-pion exchange and short-range contributions,
Phys. Rev. C 84, 054008 (2011).
arXiv:1107.0602 [nucl-th],
doi:10.1103/PhysRevC.84.054008.
[15]
H. Krebs, E. Epelbaum and U.-G. Meißner,
Nuclear axial current operators to fourth order in chiral effective field theory,
Annals Phys. 378, 317-395 (2017).
arXiv:1610.03569 [nucl-th],
doi:10.1016/j.aop.2017.01.021.
[16]
H. Krebs, E. Epelbaum and U.-G. Meißner,
Nuclear Electromagnetic Currents to Fourth Order in Chiral Effective Field Theory,
Few Body Syst. 60, no.2, 31 (2019).
arXiv:1902.06839 [nucl-th],
doi:10.1007/s00601-019-1500-5.
[19]
J. L. Friar,
Pion Exchange Contributions to the Nuclear Charge, Current, and Hamiltonian Operators,
Annals Phys. 104, 380-426 (1977),
doi:10.1016/0003-4916(77)90337-2.
[20]
J. Adam, H. Goller and H. Arenhovel,
Relativistic corrections and unitary equivalence in elastic electron deuteron scattering,
Phys. Rev. C 48, 370-378 (1993),
doi:10.1103/PhysRevC.48.370.
[21]
A. A. Filin, D. Möller, V. Baru, E. Epelbaum, H. Krebs and P. Reinert,
High-accuracy calculation of the deuteron charge and quadrupole form factors in chiral effective field theory,
Phys. Rev. C 103, no.2, 024313 (2021),
arXiv:2009.08911 [nucl-th],
doi:10.1103/PhysRevC.103.024313.
[22]
L. Girlanda, A. Kievsky, L. E. Marcucci and M. Viviani,
Unitary ambiguity of NN contact interactions and the 3N force,
Phys. Rev. C 102, 064003 (2020),
arXiv:2007.04161 [nucl-th],
doi:10.1103/PhysRevC.102.064003.
[23]
L. Girlanda, E. Filandri, A. Kievsky, L. E. Marcucci and M. Viviani,
Effect of the N3LO three-nucleon contact interaction on p-d scattering observables,
Phys. Rev. C 107, no.6, L061001 (2023),
arXiv:2302.03468 [nucl-th],
doi:10.1103/PhysRevC.107.L061001.
[24]
E. Epelbaum, H. Krebs and P. Reinert,
Semi-local Nuclear Forces From Chiral EFT: State-of-the-Art and Challenges,
in Handbook of Nuclear Physics, edited by I. Tanihata, H. Toki, and T. Kajino (2022) pp. 1–25,
arXiv:2206.07072 [nucl-th],
doi:dx.doi.org/10.1007/978-981-15-8818-1-54-1.
[25]
E. Epelbaum, H. Krebs and U.-G. Meißner,
Improved chiral nucleon-nucleon potential up to next-to-next-to-next-to-leading order,
Eur. Phys. J. A 51, no.5, 53 (2015),
arXiv:1412.0142 [nucl-th],
doi:10.1140/epja/i2015-15053-8.
[26]
R. Machleidt, P. Liu, D. R. Entem and E. Ruiz Arriola,
Renormalization of the leading-order chiral nucleon-nucleon interaction and bulk properties of nuclear matter,
Phys. Rev. C 81, 024001 (2010)arXiv:0910.3942 [nucl-th],
doi:10.1103/PhysRevC.81.024001.
[27]
P. Reinert,
Precision studies in the two-nucleon system using chiral effective
field theory,
PhD thesis, Ruhr University Bochum, 2022,
doi:10.13154/294-9501.
[28]
E. G. Kessler, Jr.,
The Deuteron Binding Energy and the Neutron Mass,
Phys. Lett. A 255, 221 (1999),
doi:10.1016/S0375-9601(99)00078-X.
[29]
J. J. de Swart, C. P. F. Terheggen and V. G. J. Stoks,
The Low-energy n p scattering parameters and the deuteron,
Contribution to the 3rd International Symposium on Dubna Deuteron 95,
arXiv:nucl-th/9509032 [nucl-th].
[30]
N. L. Rodning and L. D. Knutson,
Asymptotic D-state to S-state ratio of the deuteron,
Phys. Rev. C 41, 898-909 (1990),
doi:10.1103/PhysRevC.41.898.
[31]
J. L. Friar,
Measurability of the deuteron D state probability,
Phys. Rev. C 20, 325-330 (1979),
doi:10.1103/PhysRevC.20.325.
[34]
P. W. Lisowski, R. E. Shamu, G. F. Auchampaugh, N. S. P. King, M. S. Moore, G. L. Morgan and T. S. Singleton,
Search for resonance structure in np total cross-section below 800 MeV,
Phys. Rev. Lett. 49, 255-259 (1982),
doi:10.1103/PhysRevLett.49.255.