1. Introduction
The study of the correlations between the particles of a system is key to the understanding of its behavior, classical or quantum. Without trying to be exhaustive, one may mention the following main lines in which are involved: (a) equilibrium thermodynamics [
1,
2,
3,
4,
5,
6,
7,
8,
9,
10,
11,
12,
13]; (b) time-dependent phenomena [
12,
13,
14,
15,
16,
17,
18,
19,
20,
21,
22]; (c.1) the experimental [
23,
24,
25,
26,
27,
28,
29,
30,
31] and/or (c.2) computational [
7,
8,
9,
10,
11,
12,
13,
17,
32,
33,
34,
35,
36,
37,
38,
39,
40,
41,
42,
43,
44,
45,
46] techniques for their practical determinations; and (d) the fundamental questions related to entanglement and separability [
47,
48,
49,
50,
51,
52,
53]. Each of these lines forms a whole body of knowledge with direct/indirect practical applications, ranging from phase equilibria and stability to quantum networks and information. Furthermore, all of them taken together display useful and revealing intersection areas. In this connection, the present article will focus on statistical mechanics issues at the intersection between lines (a) and (c.2), by selecting the somewhat forgotten topic of the triplet correlations in equilibrium
quantum monatomic fluids (in 3-D space) at nonzero temperatures. This situation contrasts sharply with that of the classical fluid, for which insightful theoretical and computational developments have been available in the literature for a long time (see for instance References [
37,
38,
41,
42,
43,
44,
45]).
As a convention, the concept “monatomic fluids” can imply either actual fluids composed of atoms or model fluids composed of structureless particles (one-site). In this work, both terms atom and particle will be used indistinctly. Also, it is convenient to stress that in a broad sense the general concept of equilibrium “correlations” encompasses the usual fluid structures
in the real
r-space and the response functions
in the reciprocal Fourier
k-space (
n refers to the number of particles involved in their formulations). Both types of
static structures are connected essentially through Fourier transforms, which are far from trivial for
[
3,
7,
9,
11,
41]. One also notes that an extended use has assigned the names
correlation functions and
structure factors to the
and the
respectively. This is the convention followed in this article, although depending on the context “correlations” or “structures” will be used occasionally in the global sense.
For homogeneous and isotropic monatomic fluids, at the pair level
one deals with 2-D structural functions, whilst at the triplet level
one deals with 4-D structural functions. Contrarily to the pair functions, the triplet functions cannot be extracted from radiation scattering experiments, because the triplet contribution to the scattering intensity is negligible as compared to the rest of the contributions that shape the differential cross section [
24]. Therefore, a comprehensive quantum fluid triplet study, which goes beyond the time-honored closure-theory attempts [
3,
39,
40], entails an appropriately developed theoretical framework, which must be complemented with very powerful computational means. In this way, one can perform the extensive numerical calculations needed (the experiments here). From recent path-integral Monte Carlo plus approximate closure works by the present author [
54,
55,
56,
57,
58,
59,
60], one may feel that the time is becoming ripe to start undertaking the triplet topic. As in the classical domain the interest in this quantum task does justify by itself, if only because it makes statistical mechanics go beyond the usual pair level. But for those more application-orientated, one may point out that there are many questions which can benefit from such development (e.g., further fluid structural characterizations of the fluid-solid transition [
6,
41,
61,
62], phonon-phonon interactions in superfluids [
40], time-dependent phenomena [
16,
17,
18,
19,
20,
21,
22,
23,
33,
34] (see Reference [
24] to infer the possible connections), colloid suspension properties [
63,
64], etc.). In relation to this, it may be motivating to mention the empirical relationships found for the maxima of the equilateral correlations in
r-space of the quantum hard-sphere (QHS) fluid along the crystallization line [
59], and the surprising success of closures in capturing significant quantum fluid triplet features [
59,
60]. The reader should be aware however that the quantum fluid complexity is far greater than that of its classical fluid counterpart, as will be discussed in detail in this article.
The outline of this article is as follows.
Section 2 is devoted to giving an introductory global description of the context in which the quantum triplet topic is inserted, together with its current state of development.
Section 3 contains a condensed presentation of the pair and triplet structures in classical statistical mechanics. Why? Because, although there are radical differences between the classical and the quantum domains, both share a good deal of the general notations involved, plus a significant number of basic concepts and tools used (e.g., functional differentiation and closures). This excursion into the classical background is expected to ease the way to the quantum fluid discussion.
Section 4 reviews the basic theory of Feynman’s path integrals PI in (thermal) imaginary time [
4,
7,
9]. When complemented with computer simulations (resembling somehow the classical ones), PI can be utilized for the complete study of quantum fluid static triplets (
r-space and
k-space) in the diffraction and the bosonic exchange regime (more on the exchange issues later).
Section 5 focuses upon the higher complexity of the quantum fluid structures as revealed by PI, concentrating on the quantum diffraction regime by reason of its fundamental role. Thus, one finds that the single structural class present in the classical domain, C, composed of pairs
splits into three different structural classes, namely instantaneous ETn,
total continuous linear response TLRn,
and centroids CMn,
[
7,
9,
11]. The discussion focuses on
, although paying due attention to
, by reason of its importance when studying triplets. The parallel extension to zero-spin particles and bosonic fluids is deferred to an Appendix. To illustrate some of the main points discussed, a set of various results are included in this
Section 5. The studied systems are helium-3 under supercritical conditions, and the quantum hard-sphere fluid on its crystallization line. For helium-3 the results are related to the
r- and
k-spaces, while for quantum hard spheres are on
k-space.
Section 6 gives a summary description of the systems, and the methods used to facilitate the understanding of the reported results. The methods are path integral Monte Carlo simulations (PIMC) and a significant number of closures, such as Kirkwood superposition (KS3) [
37], Jackson-Feenberg (JF3) [
40], the intermediate AV3
(KS3+JF3)/2 [
59], Barrat-Hansen-Pastore (BHP3) [
41], and Denton-Ashcroft [
65].
Section 7 contains a series of final remarks together with directions for future research. Finally, two Appendices, I the boson fluid and II a list of acronyms and basic associated references for the reader’s convenience, close this work.
5. Theory of Equilibrium Quantum Fluid Triplet Structures under Diffraction Effects
A homogeneous and isotropic quantum monatomic fluid is considered. Exchange interactions are neglected and the Hamiltonian
Equation (1) is selected. The correlation functions and their associated structure factors can be grouped into three classes [
7,
9,
11,
84]: centroids (CMn), total continuous linear response (TLRn), and instantaneous (ETn). Each class is associated with the linear response from the fluid to an external weak field
. When use of functional calculus techniques is made, a good deal of the developments for the CMn class are parallel to that of the classical monatomic fluid, and the same may be said of the TLRn class, though to a significantly lesser extent [
7,
11,
55,
130]. The ETn case has a different nature, the functional techniques that can be applied to CMn and TLRn are inapplicable to ETn, and this latter class is to be treated separately [
3,
6,
9,
84,
135].
The grand-canonical partition function for the study of the CMn and TLRn classes incorporates the action of
and reads as follows [
11,
62]:
where the operator
is defined as
The key point here is how to deal with the exponential of the sum of two non-commutative operators, i.e.,
, which can be accomplished in the standard way via the Baker-Campbell-Hausdorff formula [
163]. For the current purposes, the density matrix elements in Equation (41) are split into
X steps and one applies the basic approximation:
which is accurate up to
terms (as occurs in the PP propagator derivation). This approximation leads to the partition function:
is diagonal in the coordinate representation). By observing Equation (43) one notes that the nonnegative density matrix elements may be given the final PI propagator form desired or needed for PI convergence reasons. In relation to this, notice that: (a) functional derivations involving
are independent of such form, (b) the linear response from the fluid is characterized by functions in the limit
and (c) the optimal value for
X can therefore be conveniently adapted. By recalling that, regardless of the propagator employed, the final optimal number of beads in PI calculations is conventionally taken as
P, the latter remark (c) is unimportant for PP and CBHSP (or PA’s in general), since
. However, it turns out to be crucial for SCVJ for which
and the physically significant beads for structural purposes are just those contained in Equation (43):
. (Note that
is the primary object that serves to define the structural ensemble averages). As stressed earlier, such significant
beads are the odd-numbered
t ones when SCVJ is fully developed and the intermediate beads completing the whole
P-set do arise, i.e.,
[
11,
84]. Consequently,
can be approximated by the general PI form [
11,
62]:
Now, it is worthwhile to point out that the action of the field expressed in Equation (44) is consistent with a thermalized
-interaction, but not with a “sudden” interaction such as that of the radiation-atom elastic collisions. On the one hand, examples of the fields in 3-D space involved in Equation (44) are: static continuous fields in
r-space and continuous fields in imaginary time, such as a gravitational field or the special case of a set of neutrons thermalized with a fluid sample composed of zero-spin particles, as is the case of fluid helium-4 [
30]. (For the latter, such continuous interaction takes place in Trotter’s limit
, which is PI treated in (44) via the finite number of beads,
P or
. The classes CMn and TLRn are associated with these types of continuous
-fields. On the other hand, examples of related fields not involved in Equation (44) correspond to the elastic scattering of radiation experiments, e.g., neutrons for fluid helium-4, X-rays for fluids helium-4 and helium-3 [
6,
28] (from a conceptual statistical mechanical point of view there is no difference between the static response functions to both radiation probes [
6]). Thus, within the framework set in subsection 2.2, the inadequacy of the static picture related to Equation (44) can be illustrated, for simplicity, by considering the interaction between a neutron and a spinless particle, which is essentially given by the neutron-nucleus Fermi’s potential (a Dirac-
) [
6,
23]: the (elastic) collision between an incoming neutron and a thermally delocalized atom localizes the atom at just a position in the sample (the “collapse” by the position measurement); afterwards the atom must delocalize again according to the fluid equilibrium state. “Viewed” from the foregoing PI approach, the neutron-atom collision should take place at the position of a given bead, which would imply the immediate disappearance of the associated necklace distribution and its ensuing reappearance [
11,
170]. This quantum phenomenon cannot therefore be analyzed via the functional derivatives of (44), which cannot cope with the disappearance/ reappearance of the atom quantum thermal packet, and a full quantum time-dependent treatment is the way to understand it [
6]. Therefore, for the class ETn, which is associated with localizing
-fields, Equation (44) is not appropriate (more about this in subsections 5.3 and 5.4). In sharp contrast, under classical conditions the radiation-atom elastic collision phenomena pose no problems to the functional derivations and their
static linear response interpretation [
5,
6,
148]. Now, a detailed discussion of the three classes of equilibrium quantum structures follows.
5.1. The PI-Centroid CMn Class
This class behaves from the structural standpoint in the classical way. This is a remarkable result that arises from the explicit consideration of a continuous weak field
of constant strength,
acting on the fluid [
19,
33,
55,
84,
130]:
Inclusion of this field leads to the general PI partition function [
130]:
where use of Equation (40) is made. For clarity reasons, it is convenient to rewrite the previous formula as:
It is straightforward to show that the functional derivatives of
are formally identical to those of the classical case contained in Equations (8)-(9). Thus, one finds up to the third order [
11,
57,
130]:
where the inhomogeneities caused by
are reflected in the fluid centroid functionals/properties
The CMn hierarchy can be obtained by iterating the procedure to higher orders. As explained earlier in the classical case, if one makes
on the right-hand sides, the isolated fluid properties arise:
where
and use is made of the total pair correlation function of centroids
Incidentally, note that the factor accompanying
may be abbreviated to
, consistently with the subtraction from
of the asymptotic behaviors
Note that Equations (49b) and (49c) can be formulated in a concise manner with the use of the spontaneous fluctuations of the centroid density (analogous with Equations (9b) and (9c)).
The formal expressions for the grand-canonical averages giving
and
are identical to those of the classical case Equations (3) and (4) (
Figure 1):
5.1.1. PI-Centroid Linear Response
By integrating (48b) and (48c) with the substitutions (49b) and (49c) on the right-hand sides, and taking the Fourier transforms, the two- and three-body centroid response functions arise as [
11,
55]:
The same dependences on the moduli of the wavevectors (e.g.,
and similar comments as those made in connection to the classical Equations (10)-(11) do apply here. In particular, the
terms are key to formulating completely
and
, and correspond to the especial situations: (a)
; (b)
and/or
, and
Explicitly, the CM3 triplet structure factor can be cast as:
where
stands for the Fourier transform of
does for that of
, etc. Given the analogies between the structural formulations for quantum centroids and classical monatomic particles, it would be a significant step forward if such parallelism could be extended to also include the OZn direct correlation function schemes for centroids. This is indeed the case as discussed in detail below [
55,
130].
5.1.2. PI-Centroid Direct Correlation Functions
There is the direct axiomatic way, consisting in checking that the OZn relationships satisfied in the classical domain by the set
(see the insightful work by Lee [
70]) are also satisfied by a set of PI-centroid functions
which is defined accordingly in the same manner (see Reference [
130] for OZ2). This procedure finds support in: (a) the formal equivalence between the PI-centroid and the classical hierarchies arising from their partition functions, and (b) the fact that the external field is switched off
when deriving the usual OZn equations [
70], which serve to analyze the structures of the isolated fluid equilibrium states. Nonetheless, it seems useful to obtain the centroid OZn equations via a free-energy argument closely related to the standard procedures [
171,
172,
173,
174,
175]. The reasoning below expands in a more comprehensive manner the early theoretical sketch reported in [
55], and hinges heavily on the external field of constant force
as the cause for the PI-centroid structures to show up. Recall that the PI-centroid (a “center-of-mass” variable) is an auxiliary mathematical object which, as such, cannot interact with external fields in an experimentally measurable direct way. Therefore, no claim can be made regarding any general validity of the centroid OZn(
) derivations under other weak external potentials
, as is usual in the general functional OZn-context of classical fluids. The following development is an adaption to PI-centroids of the procedure that can be found in Haymet
et al.’s insightful works [
173,
174,
175].
(i) The starting point is the differential of the fluid energy:
where
S is the entropy,
p the pressure,
the one-body number density distribution of actual particles, and
a (constant) variation in the field acting on such particles. The key point is that under these conditions Equation (55) can be rewritten in terms of the centroids:
because it so happens that
Equation (57) can be obtained directly by integrating the functional derivative
(see Equation (48a) and subsection 5.2). However, at this point it seems better to consider the role of the normalized thermal packet
representing a delocalized atom in the inhomogeneous fluid. The function
is a probability density necessarily normalized to unity:
where
is the position vector of the atom
, and
the position vector of the centroid
. To grasp this concept, consider the simpler example of the homogeneous Gaussian thermal packet provided by the semiclassical Feynman-Hibbs picture (GFH) [
4,
11,
143,
144]), in which a particle of mass
m in a fluid is described as a thermal packet of width
[
4]. The corresponding
may be visualized as a deformation of the Gaussian packet adopting the symmetry of the external field. The derivation of Equation (57) uses the following facts: (a) the centroid is the “center of mass” of the probability density
; (b) the actual atom density
is obviously related to the centroid density
through a smearing out operation involving the inhomogeneous
packet; and (c) the external force undergoes a constant variation
The proof goes as follows
:
(ii) Following the steps in References [
173,
174,
175] one defines an auxiliary energy via a Legendre transformation:
and a Helmholtz free energy for the fluid in the form:
(iii) A discussion of the selection of reference ideal systems in classical and quantum studies can be found in References [
172,
173,
174,
175]. This is a very important matter, for the application of functional differentiation to the corresponding excess free energy will yield a hierarchy of direct correlation functions. The direct correlation functions so obtained may or may not coincide with the usual for classical simple fluids [
70,
146,
147,
171,
172,
173,
174,
175]. In the current case of PI-centroids, which are quantum intermediate objects [
4,
11,
18,
19,
140], given the classical-like form of the derivatives of the grand canonical potential,
, with respect to the field
given in Equations (48), the selection of the ideal Boltzmann system in the field
seems the appropriate choice. By doing so, a classical-like hierarchy of PI-centroid direct correlation functions will be obtained. For the present structural purposes, focused on the usual Ornstein-Zernike framework at
this choice is going to be rewarding. Therefore, the Boltzmann-system reference [
171] yields [
55]:
and the excess free energy
reads as:
(iv) The functional differentiation of
leads to the hierarchy of the PI-centroid direct correlation functions
which, up to the third order, reads as follows:
Of particular interest is the pair function
which can be developed in the convenient form:
where attention should be drawn to the inverse derivative
which is perfectly consistent with the use of the grand canonical ensemble, for density variations independent of the external field can be made [
147].
5.1.3. OZ2 and OZ3 Frameworks
The OZ2 framework for PI-centroids arises in the usual way by applying the matrix identity [
5,
147]
Inserting Equations (48b) and (65) into Equation (66), one arrives at the inhomogeneous OZ2 equation in the field
which transforms into the conventional homogeneous version if the field is switched off
By Fourier transforming Equation (68) and combining the result with Equation (52c), the static pair structure factor for PI-centroids is:
formally identical to the classical form Equation (13a). It can be demonstrated that for the quantum fluid [
130]
a result that should not be surprising, since the fluctuations in the number of particles can be counted with the centroids. Therefore, the PI-centroid
component fixed via
formally provides an exact way to obtain the equation of state of a fluid with quantum diffraction behavior. As in the classical case,
is expected to decay effectively to zero within a short range of distances, which is indeed the case and leads to highly accurate results in
k-space [
46,
62]. In this connection, recall that direct correlation functions show generally a tail in their decays [
149], but its importance is negligible beyond a certain cut off (see References [
130,
136] and
Figure 2).
At this point, and following Reference [
41], it is worth introducing a new hierarchy
arising from the functional derivatives of the total free energy
. As before, the following formulas are limited to the third order, the iteration to higher orders being straightforward. For brevity, detailed reference to the various dependences of the functionals are omitted in what follows. At first order one finds:
where the terms containing
come from the ideal system contributions. In Equation (71) one identifies the actual chemical potential
as the algebraic sum of the external field and an intrinsic contribution, in agreement with the expected constancy of the chemical potential in the presence of the external field [
171,
172].
The interest of the
hierarchy lies in the fact that, together with the hierarchy
they allow the OZn frameworks to be formulated in compact ways [
41]. Thus, OZ2 as given in Equation (67) can be cast as:
which reduces to Equation (68) in the limit
Furthermore, via functional differentiation of Equation (74) one obtains the OZ3 framework. One of the four equivalent OZ3 equations can be derived by following closely References [
41] and [
70]. First, by taking the functional derivative with respect to the one-particle density a new identity arises:
Second, with application to
of intermediate differentiation with respect to
, and renumbering the particle labels, one arrives at the final form:
which, again, reduces to the homogeneous OZ3 equation in the limit
The triplet linear response function for PI-centroids of the homogeneous quantum fluid
arises from the Fourier transform of Equation (76) combined with the definition given in Equation (53b). Thus, by noting that:
one finds the triplet structure factor:
which is equivalent to Equation (54) but formulated in terms of direct correlation functions [
55], with all the advantages and problems associated with these objects commented earlier. In this connection, if the variations in the density are uniform, the general hierarchy
sketched in Equations (64) takes the simpler and operative Baxter’s form [
68]. In particular, the analogous to Equations (14) read as:
5.1.4. Some Additional Relationships and Further Remarks
For the hierarchy
the partition function
leads to all the relationships that are applicable in the classical context [
70]. A straightforward calculation yields:
which at zero field gives the corresponding centroid equations for the isolated quantum fluid. Completing with the first order derivative
, the first three of the related equations read as:
Besides, note the classical-like additional identities that hold in the PI-centroid context:
where
is the in-the-field PI canonical partition function, and the relations to the number fluctuations are clearly displayed. These number fluctuations refer to the in-the-field fluid; the formulas on the right-hand sides remain the same when
but obviously the meaning is not the same.
Finally, several remarks are in order. It is worthwhile to point out that the formal OZn derivations applied in the classical domain and related to the bijective connection external field
equilibrium density [
171,
172,
175,
176,
177], hold here for
. This is guaranteed by the form of the quantum partition function Equation (44) which is isomorphic to its classical counterpart Equation (2). Another point is that there is no defined interparticle potential
between PI centroids (compare to the variational approximations arising from the PI formalism [
4,
140,
141,
142], which yield versions of such a potential), but this is not relevant to this discussion [
171]. In this regard, nor is the fact that under the same conditions two different densities are included in Equation (57), namely the true particle density and the centroid density, because: (a) one is the actual and the other is an auxiliary quantity, (b) each corresponds to a distinct object, and (c) both are intimately related. (The situation is analogous to a distribution of mass and the consideration of its center of gravity). To bring this subsection to a close, note that the applicability of the classical OZn numerical schemes to the quantum CMn class is perfectly possible and rewarding: (i) at the pair level, OZ2(BDH+BHw) is a powerful way to fix the equations of state of quantum fluids [
46,
62,
130,
136,
137]; and (ii) at the triplet level, first OZ3 applications were reported in References [
55,
56,
58].
5.2. The PI Total Continuous Linear Response TLRn Class
For this class the fluid undergoes a general external weak field
which is continuous either spatially or in imaginary time (no magnetic interactions are considered). The PI partition function can be cast as [
7,
11,
57]:
The functional derivatives for this model now can be cast as:
In the foregoing formulas a vector
stands for the position vector of a structurally significant bead in the grand-canonical sample (i.e.,
is any of the
Recall that for structural evaluations each canonical ensemble has
significant particles, and that
X is held fixed throughout the grand canonical ensemble. Therefore, by following the standard procedures [
2,
11,
57] one sets the definitions:
and obtains the relationships:
where the number fluctuations refer to the in-the-field fluid. The formulas on the right-hand sides remain the same when
although the meaning is therefore not the same.
The linear response cases in this case arise by making
on the right-hand sides of Equations (82). In
r-space this can be summarized in the following equations [
57]:
where it is important to note the role played by
X in Equations (84). Equations (84b) and (84c) can be formulated in concise forms with Equations (82) and the spontaneous fluctuations of the bead density at the two- and three-body levels respectively.
For notational simplicity, by
arranging correlatively the
NX beads present in each canonical ensemble
the pair
and triplet
correlation functions are defined by the ensemble averages [
57]:
Note that the correlations between beads include both: equal and different imaginary times. The foregoing averages (85a)-(85b) can be split into actual atom components: one-particle, two-particle, and three-particle. In this regard, (a) for
: the bead-bead pair correlations can correspond to beads in the same necklace (one-atom self-correlations
) or to beads in different necklaces (two-atom correlations
) [11}:
and (b) for
the bead-bead-bead triplet correlations can correspond to three beads in the same necklace (one-atom self correlations
), two beads in the same necklace and the third in another necklace (mixture of one-atom self-correlations and two-atom correlations
), and the three beads in different necklaces (three-atom correlations
. At the triplet level this can be cast as [
11,
57]:
where for clarity reasons the vector notation and vertical bars to separate necklaces in the
cases are used.
The pair and triplet static structure factors arise from Equations (84b)-(84c) via the Fourier transform process explained earlier. They can be cast as [
11,
55,
56,
57]:
which in expanded form can be written as:
More information about these structural functions in the
r- and the
k- spaces can be found in works by this author [
11,
46,
54,
55,
56,
57,
62].
The fixing of the
and
spatial structures can be accomplished with PIMC or PIMD simulations. Note that their computations scale as
and
respectively, which makes
a demanding task. To date these calculations have focused on full evaluations of
[
11,
46,
62], but only on restricted evaluations of
covering equilateral and isosceles features of the three-particle part
of quantum hard spheres [
54]. As for the structure factors
and
the PIMC/PIMD situation regarding the
X-scaling is similar but further aggravated by the necessity to scan sets of commensurate wavevectors and, also, by the drawbacks associated with the low-
k regions. Again, a significant part of
, not very close to
can be determined with reasonable PI simulation effort [
46]. Nonetheless, the simulations of
over a significant range of wavevectors, excluding obviously the unattainable
and/or
, and
, remain today daunting (at least for most researchers).
OZ2 and OZ3 Frameworks
In view of the present-day PI-simulation difficulties in obtaining even restricted descriptions of the structure factors TLR2 and (above all) TLR3, one might think of resorting to direct correlation functions. However, the delocalization of quantum particles brings about several complications if one tries to construct OZn frameworks by following classical-like ways. A couple of examples will suffice to illustrate this knotty issue. Firstly, just at the pair level, one can consider Haymet
et al.’s approach [
173,
174,
175]. Here, the pair part of the TLR2 case Equation (87a), which depends only on
and is normalized to unity for large wavevectors (see below), was the object of interest as applied to freezing in fluid helium-4 [
175]. Broadly speaking, some of the main features are: (a) the ideal system was defined as the Feynman ideal system; (b) the direct use of inverse derivatives
led to a need for effective particle masses in the ideal system to deal with the self correlations, though this measure could not be implemented under diminishing temperatures; and (c) in the end, a generalized OZ2 framework was obtained, but “effective” structure factors had to be defined, and unnatural amplifications in that pair part of TLR2 were detected for large wavevectors
k. Secondly, another generalized OZ2 framework is that put forward by Shinoda
et al. [
178,
179], in which the time-honored molecular approach RISM [
7,
71] was adapted to tackling the PI-model of a fluid with quantum behavior, as if it were a classical molecular fluid (RISM = reference interaction site model). This approach also focused on the same pair part of TLR2 mentioned above, and self interactions did appear when working the related OZ2 scheme arising from RISM (i.e., a given particle “interacts” with itself at different imaginary times). Accordingly, further approaches to cope with this fact of the model were to be taken.
At this juncture, the following two observations may be useful to help grasp the difficulties in this context. (i) First, a simple exercise. By assuming that for the TLR2 unrestricted bead-bead correlations one might define a direct correlation function in the conventional manner [
5]:
it is straightforward to check the strange
behavior that application of an identity analogous to (66) would yield with increasing
X As in the two foregoing approaches, one can rederive the classical OZ2 equation for
but the expression for the Trotter limit
turns out to be meaningless. (ii) Second, a mathematical fact. Note that most of the OZ2 frameworks (e.g., Percus-Yevick, hypernetted chain, and their variants [
69,
71]) are formulated with reference to interatomic potentials
The point is that the potential energy operators
are diagonal in the coordinate representation, which means that beads do interact via
if and only if they belong to different necklaces and are at the same imaginary time step (e.g., as shown in Equations (36a), (37a), and (38)). Hence, there cannot be any particle self interactions in a PI context.
The approximation proposed by the present author to deal with
consists in separating the intra-necklace bead-bead correlations from the inter-necklace bead-bead correlations and applying directly the classical OZ2 framework to the latter (the pair part mentioned above) [
180]. This (lucky) choice found some support by the applicability of this procedure to the GFH picture of fluids with (weak) quantum behavior [
11,
143,
144,
180]. This approach can be summarized as follows:
where
is defined by the first two terms in Equation (89a), being a sort of “form factor” for the one-particle thermal packet, and
adopts the well-known classical form associated with the overall inter-necklace correlations
The self correlations do not contribute to the isothermal compressibility of the fluid (i.e., the
component), which can also be formulated in the usual way only with
[
130]. Application to
of the standard OZ2 Equation (12), which defines its associated
yields after Fourier transforming:
where one notes that
admits approximate analytical representations [
131,
180], a very useful one being that derived within the GFH picture [
131]
One way or another, quantum approximations that involve classical
schemes provide this context with reasonable alternatives to handle
k-space difficulties [
46,
173,
178,
180], although their corresponding ranges of validity are to be analyzed
a posteriori. Surprisingly, for the approximation given by Equation (90a), such range turns out to exceed expectations, since even under very strong diffraction effects the comparisons with experiment and PI-simulations are excellent. In this respect, at
for supercritical helium-3 and for normal liquid helium-4 at and near SVP (saturated vapor pressure) conditions, the TLR2 results based on OZ2(BDH+BHw) were found to be almost indistinguishable from those obtained via PIMC, and/or very close to experiment [
30,
46,
170,
181]. Under diffraction effects and at the pair level, the conclusion when conducting OZ2 treatments is clear: the intra- and inter-necklace bead-bead correlations can be separated safely for most practical purposes. Two more observations: (a) with increasing quantum effects, the very low-
k region may be affected by the OZ2 approximations used (see subsection 5.4); and (b) when exchange interactions are not negligible, this sort of bead-bead separation cannot be applied.
So far, to this author’s knowledge, no related OZ3 schemes have been reported for TLR3. The intricacy of these triplet features is certainly greater than at the pair level. In relation to this, one notes that the straightforward procedure followed in the pair linear response structures, Equations (86a)-(89)-(90), cannot be used in the functions leading to , as shown in Equations (86b) and (87c). Accordingly, whether this type of OZ ideas might be fruitful when studying the triplet remains to be investigated. The use of approximations, extending those in Equations (90), for treating the terms displayed in Equation (86b) will be necessary.
5.3. The PI Instantaneous ETn Class
This last class is not amenable to being treated with functional derivatives involving an external weak field, because of the collapse of the thermal packet under an instantaneous localization process. To obtain the definitions of the related pair correlation function
and its associated structure factor
which are in many ways analogues to their classical counterparts
the standard quantum reasoning is based on the study of: (a) neutron scattering, with the time-dependent treatment using Born’s approximation and Fermi’s potential, plus the elastic sum rule applied to the dynamic structure factor
[
6,
23,
29]; or (b) the elastic scattering of X-rays, plus the consideration only of the pair coherent part associated with the nuclei (i.e., the single-atom actual form factor due to the electrons is dealt with separately) [
6,
28]. For the diffraction quantum effects the final expressions are [
3,
6,
9,
11]:
where the ensemble averages refer to
Equation (29), both being formally equivalent to the classical definitions
. The latter remark implies that both quantum averages reduce to the usual formulas in the classical limit. Explicitly, the pair correlations are defined via the ensemble average:
and the PI approach gives:
or in compact form
where the selection of the structurally significant (and equivalent) equal
-time beads is to be highlighted; this is the distinctive trait of the ETn class. Likewise, the instantaneous structure factor can be calculated within PI as [
9,
11]:
It is a simple matter to extend the formulation for the instantaneous triplets in
r-space, the PI average being (
Figure 3):
The spontaneous fluctuations in the density at the pair and the triplet levels can be expressed by formulas, involving
Equation (30), analogous to the classical Equations (9b)-(9c):
and, therefore, the actual and the PI definitions of the pair and triplet instantaneous structure factors can be written as:
Once again, this definition of the triplet structure factor is very convenient, since it reduces to the classical formulation when neglecting quantum effects [
3]. For completeness, note that within PI one finds the explicit triplet formulas [
55]:
Some related remarks are the following. (a) From the simulation standpoint: note that the ETn structural computations scale as X, which make them more expensive than those of the classical case (but far less expensive than those of TLRn which scale as as shown in Equation (87b). (b) The development of Equation (100) is formally the same as that of the classical case or, equivalently, that of the PI centroids given in Equation (54), where one only has to substitute the CMn quantities for the ETn ones. (c) As stressed earlier, there is a full formal equivalence among the Cn, CMn, and ETn basic formulations of the isolated fluid quantities in r-space and k-space
5.3.1. OZ2 and OZ3 Frameworks
At this stage of the presentation, it is clear that no exact OZn frameworks are available for analyzing the structural functions of the ETn class. However, as occurred in the TLRn class, the adoption of classical schemes/ideas, either as such [
3] or modified [
69], seems unavoidable if one tries to use the device of direct correlation functions. In this regard, and again based on the applicability of the classical-OZn to the PI-based variational pictures [
11,
143,
144,
145,
180], there was nothing to lose by trying the direct use of the classical OZ2 and OZ3 schemes. It goes without saying that the OZ2(BDH+BHw) results were beyond expectations. Thus, under very strong quantum diffraction effects (e.g., helium-4 and helium-3 fluids at
), the overall agreement with experiment and PI simulation was excellent [
28,
29,
46,
55,
170,
181], the very low-
k region being the most sensitive to this approximation (see subsection 5.4). As regards OZ3, applications of the classical scheme have been made to helium-3 [
55,
60], the quantum hard sphere fluid (bare of with Yukawa attractions) [
56], and liquid parahydrogen [
57,
58], but no full tests of validity have been conducted.
The two basic equations here are the adaptations of the classical Equations (13) and (15), i.e., formally the same as in CMn, although now they are approximations:
where the direct correlation functions
and
are introduced as arising from the standard OZ2 and OZ3 classical procedures (
Figure 4 and
Figure 5).
5.3.2. An ETn Functional Digression
Before going any further, it is worthwhile to remark that there is a functional approach to the ET2 correlation function [
6,
135]. The point is that, by focusing on the variations of the partition function
with respect to the variations in the interparticle potential
one obtains an expression in which the correlation function
shows up. The general usefulness of this development seems scarce because there is no response from the fluid to an external field. However, for completeness, it is noticeable that one can obtain useful relations at the pair-level involving both the ET2 and the CM2 structures [
135]. To simplify the following discussion, only the PP propagator is used in what follows.
The PP grand canonical partition function for the homogeneous and isotropic fluid can be cast as [
7,
90]:
A straightforward calculation yields the related first functional derivative:
which is formally identical to the classical or to the PI-ET2 counterpart in the canonical ensemble that can be found in [
6] and [
135], respectively.
As regards the extensions to higher orders, the situation is rather inconclusive. To get a feeling of the drawbacks posed by this type of formulation, it seems instructive to consider the development of
:
the main problem being that different imaginary times are mixed. One might even think of extracting artificially equal-imaginary time quantities by “tinkering” with the
t-sums, but these manipulations are of no interest to ET correlations. Furthermore, beyond the pair level expressed in Equation (103), these derivatives become increasingly entangled. Although the latter might be related to spontaneous effects within the quantum fluid, they seem to be of no use for the current study of static quantum fluid structures.
5.4. A Joint Consideration of CMn, TLRn, and ETn
(i) Apart from the general consideration that the three quantum classes of linear response functions are directly related to different forms of density-correlation functions in the isolated fluid (as in the classical case), there are some properties exactly shared by the three classes. These properties obviously belong to the isolated quantum fluid and can be identified at zero-field, . A most interesting case arises for the structure factors at zero momentum transfer(s) , because the distinct response functions of the (isolated) fluid must take the same value regardless of the pair, or the triplet, structure considered.
At the pair level there is the extended compressibility theorem [
130] that states:
It is worth remarking that the foregoing exact result can be distorted when classical-like OZ2 calculations with ET2 and LR2 are carried out. On the one hand, the discrepancies from the OZ2-CM2 exact
estimate remain in general controlled and do not alter the overall representation of the
k-space structures determined. On the other hand, the importance of the disagreement may be a matter to analyze in certain cases (e.g., near changes of phase), as both the density and the temperature play roles in this issue [
46,
62,
130,
136,
137,
180].
The generalization of Equation (105) to the triplet level, involving the double-zero momentum transfer, reads as [
57]:
By taking advantage of the OZn formulation for centroids, Equation (106) can be written in terms of
Using Equations (69), (77c) and (78b), one finds
which in view of Equations (105) and (106) leads to the following result valid irrespective of the response class [
57]:
The quantity
is a quantum fluid property related to higher-order fluctuations in the number of the atoms (particles), and it is a simple matter to prove that:
The trivial connections of the zero-field Equations (105) and (108b) with the functional Equations (80d)-(80e) and (83f)-(83g), at zero field, are to be noticed (ET and CM are completely parallel in this regard).
Also, for the single-zero momentum transfer one finds the exact CM relationship [
57]
whilst for ET an analogous equation can also be written, but owing to the OZn properties utilized it is an approximation:
It seems worthwhile to explore in future work the possibility of obtaining an equation for
based on the connection between the instantaneous
and centroid
pair correlation functions [
135]. If successful, this could help to solve the ET problems for low-
k wavenumbers, generalizing in this way the exact result connecting the properties of the centroid and the delocalized particle within the GFH picture [
131,
180]. So far, no further formulas based on direct correlation functions for TLR3 have been derived.
(ii) The symmetry properties in
r-space and in
k-space of the structures CMn, TLRn, and ETn, are translations of the classical ones seen in subsection 3.3. CMn and ETn behave in the classical way regarding the limits of increasing/decreasing distances between atoms and, also, in the limit of increasing wavenumbers (see also Figures 1, 2b, and 3). However, the class TLRn (unrestricted bead-bead-… correlations) presents some striking differences. In relation to this, suffice it to mention that for the TLR2 functions one finds in
r-space: (a)
at
increases with the
P discretization, because of its one-atom component
, and tends to unity for long distances due to the two-atom component
; (b)
for long distances, and
does not necessarily vanish when two atoms come close together (it may even be nonzero at
); (c)
for large wavenumbers. The magnitudes of these discrepancies from the standard classical behaviors depend on the intensity of the quantum effects, and the reader is referred to [
46,
54,
55,
62,
170,
180] for related results. Consequently, one may expect an interesting variety of patterns when studying TLR3 functions.
(iii) Another set of exact PI relationships can be obtained [
54] by following the classical developments [
26,
150,
151,
157] given in Equations (27) and (28). It is easy to establish that for the two-point densities (referring to actual two-particle correlations) given by:
one obtains the classical-like relationships:
The latter formula leads to the classical-like equations:
where the functions
stand for:
and, for consistency,
corresponds to its associated pair structure
(recall that in practice the
values may show deviations from the exact Equation (105)).
Note that the total continuous linear response TLR is not contained in the LR relationships, since the self-correlations do not fit into this scheme. Equations (113) arise only from the pure actual two- and three-atom correlations.
For the purposes of this work, it is also worthwhile to remark that Equation (113b) can be transformed and split into two parts
as
:
where
Given that
is just a pair quantity that can be fixed with high accuracy, whilst
is a quantity that depends explicitly on
the use of these two quantities yields consistency tests of possible closures for
. In this connection, one uses KS3 as a reference by expressing the triplet closure in the form:
By substituting a closure to be tested in Equation (113f), two comparison tests are between: (a) and , and (b) the values of the density derivative determined with the computed pair results and the corresponding estimates arising from the closure.
In relation to this, for completeness, the three basic closures KS3 Equation (22), JF3 Equation (23), and AV3 Equation (24), lead to the following expressions for their quantities in Equation (113d) (
Figure 6):
(iv) Furthermore, as regards OZn schemes, application of the classical and quantum reasonings [
38,
39,
40,
41,
59,
60,
72,
125,
126,
127] is a possible and rewarding way to obtain information on the quantum triplet CM3 and ET3 structures. Recall that the CM2 calculations have, in the end, an exact OZ2 framework, but that under any other circumstances approximations are involved. Despite the latter remark, the usual PIMC/PIMD (“raw”) pair radial structures CM2, ET2, LR2, arising from the canonical simulations, can be: (a) utilized to carry out OZ2 studies of the asymptotic decay properties [
183,
184], and (b) improved using OZ2-based treatments and corrections (e.g., the BDH+BHw analysis considered in
Section 3.4). By doing so, grand-canonical approximations for these pair functions can be determined, which may be viewed as better intermediate quantities for an intensive use of direct correlation functions in triplet calculations via closures. In this connection, although for
r-space triplets the PI simulations in the canonical ensemble are sufficient to give very useful information, note that for
k-space triplets there are no complete OZ3 tests regarding the (density-temperature) ranges of validity. The huge PI computational load involved, covering significant ranges of the two wavevectors and physical conditions, and the variety of closures for triplets are reasons for explaining this lack of knowledge (
Figure 4,
Figure 5 and
Figure 7).
(v) Finally, it is worthwhile to write the connections (“sum rules”) of the dynamic structure factor
which can be obtained through inelastic neutron diffraction experiments [
23], with the PI quantities
and
where
stands for the angular frequencies of the neutron radiation.
arises as a density-density relaxation function [
23] (Equation (87a)). The latter expressions have been utilized in fluid helium-4 studies (e.g., see References [
29,
30,
170,
178]).
6. Systems Studied in This Work, Computational Details and Related Observations
The main structural concepts reviewed in the previous Sections are illustrated in this work with further numerical applications to supercritical helium-3, and the QHS fluid on the crystallization line, thus expanding the scope of previous works on these key systems [
54,
55,
59,
60]. For helium-3 the focus is on pair and triplet structures in
r-space and triplet structures in
k-space, whereas for quantum hard spheres is on triplets in
k-space. The computational techniques employed are PIMC simulations in the canonical ensemble
(or better
complemented with OZ2 and OZ3 treatments, by using the closures and relationships addressed in this work. The simulations utilize sample sizes
inside a central cubic box of side
L, with the usual periodic boundary conditions, and the Metropolis sampling scheme involving the normal-mode algorithm (acceptance ratio 50%) [
74,
145]. One PIMC kpass is defined as
attempted bead moves, and one PIMC Mpass is
kpasses. The computed structural classes are the centroid (CMn) and the instantaneous (ETn) at the pair and triplet levels, the latter being focused on some equilateral and isosceles features. Representative results are contained in the Supplementary Material.
The state points studied are listed below:
(i) For helium-3: (
;
, (
;
(
;
The pair interatomic potential selected is Janzen-Aziz’s SAPT2 [
106], and PIMC calculations involve the SCVJ propagator with
[
84]. Information obtained in Reference [
46] on auxiliary state points about the state point at
:
is also employed.
(ii) For the QHS fluid [
59,
62,
137] only
k-space results for one state point are reported,
Complementary structures for two state points about the latter, by varying
at constant
, are also obtained in the present work; these variations are:
, which for a system with
amount to
. (Other pilot applications along the crystallization line, at
[
62,
79], have also been carried out and will be commented later). PIMC calculations involve the CBHSP propagator [
81]. Recall that there is a unique way to characterize a state point in the QHS system by using reduced units [
59,
62]: i.e., the length unit is the hard-sphere diameter,
.
PIMC canonical simulations focus on the instantaneous and centroid pair
and/or triplet
correlation functions. The sample sizes
employed are as follows: (a) for
and
in helium-3,
at
,
at
, and
at
[
46]; (b) for
in the QHS fluid,
at the auxiliary state points analyzed, and the same at the target state point
taking its pair information from calculations in Reference [
59].
The pair structures in
r-space are computed in the usual way utilizing histograms [
13], the width of the bins being
Most of the “raw” pair canonical structures are subsequently OZ2-analyzed and improved in the form already described [
46,
136]: application of the OZ2 Baxter-Dixon-Hutchinson’s (BDH) procedure [
72,
125] plus Baumketner-Hiwatari (BHw) corrections [
127]. By doing so, access to every
k-space related property at the pair level is obtained with high accuracy (
Figure 2).
As regards the MC/MD calculations of
triplets, they present a number of nontrivial subtleties, as discussed by Tanaka and Fukui [
42], Baranyai and Evans [
43,
44], and Barrat
et al. [
41]. In the current PIMC applications, the method followed in
r-space is based on References [
42,
43,
44], which for the study of quantum fluids was firstly used in Reference [
54]. The interested reader is referred to the latter reference for specific details. To facilitate the understanding of the reported results, however, suffice to say that triplet correlations are determined via:
where
is the number of times that mutual distance triplets lie within the ranges
and
(again:
the previous notation should be obvious:
In addition, to avoid redundancies in the counting, the conservative (and less time-consuming) condition
is imposed [
42]. The run lengths of the PIMC computations for helium-3 in
r-space are as follows: in between 1233-1300 kpasses for fixing the pair structures
and
gathering statistics every 8000-12500 accepted configurations; and in between 2470-2560 kpasses for fixing the triplet structures
and
, where
is included, gathering statistics every 8000-10000 accepted configurations. Statistical errors (one-standard deviation) are fixed with block subaverages and remain very small: at the main peaks they are below 0.8% (see main details in
Figure 1 and
Figure 3).
The calculations of the triplet structures in
k-space turn out to be much more involved, and a few remarks are to be made. The PIMC simulations involve the sample sizes
: (a)
for helium-3 at
(
Figure 5); and (b)
at the QHS fluid state point
(
Figure 7). These calculations are focused only on the equilateral components
instantaneous for helium-3, and instantaneous and centroid for the QHS state point. Twelve sets of
vectors commensurate with the simulation box, are scanned; each set is composed of 8 pairs such that:
the vectors being at an angle
and with the constraint
[
58]. This implies that for helium-3 at
the resulting moduli (wavenumbers) are in
, whilst for the QHS fluid state point, the moduli are in
Also, the sampling is more focused on the low-
k wavenumbers and the regions of the main peaks. PIMC run lengths are as follows: for helium-3 in between 30 and 120 Mpasses, gathering statistics every 5000 accepted configurations; and for the QHS fluid state point in between 6.5 and 60 Mpasses, gathering statistics every 8000 accepted configurations. Statistical errors (one-standard deviation) are fixed with block subaverages and remain controlled: at the main amplitudes they are: 1.8% for the instantaneous case in helium-3;
for the centroid case and
for the instantaneous case in the QHS fluid (see main details in
Figure 5 and
Figure 7)). As a complementary result, the (mean) imaginary components, which must be identically zero [
41], at the main amplitudes turn out to be:
for the instantaneous case at the helium-3 state point;
for the centroid case and
for the instantaneous case at the QHS fluid state point. Clearly, although the QHS sampling is highly informative, it must be improved. In addition, all these results give an idea of the slowness of this sort of calculations when dealing with fluids with quantum behavior. This drawback becomes more acute if the fluid state point investigated belongs to the actual crystallization line of the potential model selected to describe the fluid. In this case, the
sample size cannot be small (i.e., about one hundred particles, as the used in helium-3): the “flipping” back and forth in the simulation between the fluid and solid branches [
13] must be avoided. This is the case in the current QHS fluid application. After monitoring such an effect, it has not been observed with
but it is rather conspicuous with
. This monitoring has involved the configurational centroid CM2 structure factor
defined as normalized to unity at its crystal-structure maxima [
137]. For the current QHS study with
the latter configurational quantity remains below the maximal fluid value estimate
[
61].
Closure calculations reported here for triplets (instantaneous and centroid) in
r-space are carried out utilizing KS3, JF3 and AV3, whereas in
k-space are with BHP3 [
41,
55,
56] and DAS3 [
65]. These calculations necessitate the corresponding pair structures,
and
. Detailed descriptions of the related procedures can be found in works by this author [
54,
55,
56,
57,
58,
59,
60], although for the reader to better grasp the reported results it may be worthwhile to mention the following facts.
(i) JF3 and BHP3 share a common type of integral [
39,
41,
156], which in each case is utilized in this work as follows:
- BHP3 (
k-space) [
55]
where the
stand for the standard Legendre polynomials, and the upper limit
is utilized. (In the limit
the
-expansions are exact [
41,
156]).
(ii) For the closure analyses at constant temperature, the numerical density-derivatives of the pair
and the direct correlation functions,
and
are evaluated with the centered algorithms: Stirling’s (two-point) and Richardson’s (four-point) [
182] (Figures 5 to 7). By denoting a generic radial pair function by
, the previous algorithms can be written as:
(iii) As regards JF3 and AV3, the following observations are worth making. Comparison with PIMC results for the equilateral components of the helium-3 triplet structure factors
indicates that (
Figure 5 and
Table 1): (a) JF3 cannot capture the negative behavior within the low-
k region and its usefulness reduces to the large-
k wavenumbers, since it gives the asymptotic behavior of the structure factor; and (b) AV3 as reported in Reference [
60] yields lower values for both the negative depth in the low-
k region and the height of the main amplitude. In view of the results in
Figure 6, it is highly likely that AV3 can be improved using Abe’s developments [
38] and gives better triplet results in both
r-space and
k-space.
(iv) BHP3 minimizations for ET3 helium-3 at the state point at
were reported in [
60], where use of conjugate gradients [
185] was made, as explained elsewhere [
55,
56]. Discretizations of the
functions into 7001 points covered the distance interval
. Convergence as assessed via
was rapid and without any problems, the latter quotient reaching values
in a few hundred iterations. The explored ranges of wavenumbers for
have been extended,
for checking the properties in
k-space as
k increases: JF3 appears as the limit with increasing
k wavenumbers, and the results are consistent.
Figure 4 and
Figure 5 and
Table 1 collect representative results for this system. As seen, for the equilateral components, BHP3 [
60] gives a nice representation of the PIMC values, particularly in capturing the negative behavior for
Nevertheless, some pilot BHP3 minimizations for the QHS fluid (CM3 and ET3) along the crystallization line [
79,
137] have turned out to be affected from convergence problems. These latter results are therefore far from conclusive from a physical standpoint and only a brief comment follows. The key observation is that the convergence in this case (tested with both gradients and conjugate gradients) appears to be strongly dependent on the value of the norms
in such a way that the larger this value is, the slower and more physically doubtful the convergence path turns out to be. These norm values are clearly associated with the depths at the origin of the pair direct correlation function
. To get a better feeling of this situation some data related to the pair direct correlation functions analyzed in this work are worth quoting here:
(i) For helium-3 at (
the five instantaneous values
, corresponding to the five state points involved, are in
[
60], and no BHP3 problems with the convergence were observed.
(ii) For the QHS fluid at : the three instantaneous values involved at the origin of the interparticle distances are in while the three centroid values are in
(iii) The reference values given by the squared norms turn out to be as follows:
- -
For helium-3, using Richardson four-point derivative the instantaneous reference is .
- -
For the QHS fluid along the crystallization line [
79,
137], covering the conditions mentioned above compatible with
and using Richardson four-point derivative
, one finds values for the instantaneous reference in
and for the centroid reference in
As seen, there is a significant BHP3 difference between the foregoing results for helium-3 and the QHS fluid. Note that the comparison is better made using the actual values in
for distances, since the modelling via hard spheres for helium-3 is not consistent [
55] for the current structural purposes. In relation to this, the distinct features of the interparticle potentials
involved are to be remarked: for helium-3, continuous for
and weakly attractive at the minimum
[
106], whilst for QHS there is an infinite discontinuity at
. These potentials are obviously connected to the shapes of the pair direct correlation functions, and hence to their density derivatives. Consequently, although applications of BHP3 are clearly dependent on the latter derivatives by reason of their construction Equation (25b), in intricate regions such as the freezing transition this dependence becomes extremely sensitive to the quality of the numerical derivatives. (Recall the basic density increments
about the target state points:
for helium-3, and
for QHS). Therefore, a very careful fixing of the QHS intermediate
and
by diminishing
for increasing the accuracy in the density derivative calculations, could fix this BHP3 situation. However, apart from the computational overload, possibly including an augmented precision in the calculations, one can still foresee some difficulties. Firstly, the slow minimization of a large quantity
which, together with the fine details of the density derivatives of the pair direct correlation function involved, might misdirect the convergence path and lead to unphysical solutions. Secondly, although OZ2 for centroids is formally exact, OZ2 is an approximation for the instantaneous case, which means for the latter that extra problems can be encountered in the freezing region when applying simple corrections such as BHw. Furthermore, there is the question of the complete calculation of the triplet structure factors by including its general components. Obviously, detailed work to clarify the ranges of applicability of BHP3 is needed.
(v) On the contrary the current DAS3 calculations of
are directly worked out in
k-space and, once the necessary pair information
is available, they are straightforward. Apart from the numerical derivative problems, no intrinsic procedural problems appear. As shown in
Figure 5 and
Figure 7 and
Table 1, DAS3 also gives good representations of the equilateral PIMC results, albeit apparently not so good as BHP3 in the instantaneous low-
k region
These applications use the two-point derivative Equation (118a) for the QHS-fluid CM3 and ET3 evaluations, and the four-point derivative Equation (118b) for the helium-3 ET3 evaluations. For this closure the underlying OZ2 problems with the instantaneous class also exist, and its basic assumptions Equations (26) remain to be fully tested. Again, detailed work to clarify the ranges of applicability of DAS3 is needed.
7. Final Remarks and Future Directions
The investigation of the static triplet structures in quantum condensed matter (fluids and solids) via path integrals is a promising and challenging avenue. Although it may be said that an initial background to this topic is already established, key theoretical and computational aspects in both the real (
r-space) and the reciprocal Fourier (
k-space) remain to be explored and understood. The theoretical issues should focus on: (a) expanding further the limits set by the usual pair-level approach (solid phases [
59]), and (b) the treatment of systems composed of nonzero-spin particles, particularly fermionic systems. The computational issues should mainly address the questions of cost-effectiveness posed by the required accuracy when determining the variety of different quantum triplet structures (dimensionality
4-D), with a view to increasing the simulation
sample size and, also, the possible integration of proper two- and three-body potentials (e.g., [
107,
108]) in the structure computational schemes. In this regard, exascale computations, when widely available, could alleviate the effort to be made. (Just as a high hope, if quantum computers became a reality in practice some day, the quantum triplet problems might provide an excellent test bed). All these developments can provide very useful guidance for materials research and future higher-order structural studies.
Some specific theoretical goals within the path integral framework can be the following. (i) The consideration of general external fields including magnetic interactions and/or situations involving more than one spin state in monatomic fluids; the latter can give rise to a special set of treatments for fermionic fluids when studied with the insightful Wigner path integral schemes [
98,
99]. (ii) A formulation of the actual triplet structures in molecular fluids (e.g., hydrogen fluids) that, going beyond the overall one-site picture [
57,
58], gives full meaning to the direct atom-atom constructions [
11], or yields more global molecular approaches.
As regards the specific computational goals, one can mention the following pending path integral tasks in homogeneous and isotropic monatomic fluids (4-D problems). In the diffraction regime: (iii) the determination of the total continuous linear response functions TLR3, which remain virtually untouched [
54]; and (iv) the completion of the ET3 and CM3 structures by fixing their behaviors over significant ranges of the three significant variables which characterize them (i.e, three independent interparticle distances, or two wavenumbers and an angle). In the quantum exchange regimes, there is nothing yet related to calculations of triplet structures that has been undertaken and work on this should be welcome.
As complements to the exact path integral calculations of triplets, one may try a variety of closures. These closures are not as efficient and clear-cut defined as those at the pair level. Despite this remark, there are some results indicating that, when studying the instantaneous and centroid
and
functions within the quantum diffraction regime, closures may be a great source of information. Triplets
in
r-space have been discussed at length elsewhere [
54,
55,
56,
57,
59,
60], and suffice it to say that the results obtained through the AV3 closure Equation (24) [
59,
60] point to the possibility of achieving significant improvements based on Abe’s expansion [
38]. In this regard, it is expected that the negative effects of increasing densities on the long-range isosceles correlations could be fixed to a significant extent, which may reflect in better
k-space results. Nonetheless, closure
results for quantum triplets in
k-space remain comparatively an unknown issue. For the time being, the closures BHP3 Equations (25) and DAS3 Equations (26) give results
equilateral components) that make them worth exploring in depth. They may capture traits of the sensitive low-
k (negative) region of the equilateral ET3 and/or CM3 structure factors and, also, the behavior of the important main amplitude region. In connection with the low-
k region, BHP3 seems better adapted than DAS3. However, and based on pilot calculations, BHP3 variational calculations seem less appropriate for investigating intricate regions, such as the quantum hard-sphere crystallization line, whereas DAS3 calculations appear here as a more robust option for both ET3 and CM3. For large-
k wavenumbers the basic JF3 closure (Equation (15) with
, which is part of both closures, dominates the triplet
behavior.
The
k-space closure results obtained so far are just a beginning (e.g., [
58], and
Figure 5 and
Figure 7 herein), and more work is needed to substantiate their relevance, particularly regarding the non-equilateral components. Searching for the limits of applicability of closures, even for qualitative purposes related to the CM3 and ET3 cases in the diffraction regime, seems a rewarding task. The case TLR3 does not seem amenable to closure studies because of self correlations, although there is nothing against trying closure treatments of the pure triplet-atom part ([
54] and Equations (113)).