What Is Liquid Paper

You might also like

Download as pdf or txt
Download as pdf or txt
You are on page 1of 6

What is liquid, from Noether’s perspective?

Florian Sammüller,1 Sophie Hermann,1 Daniel de las Heras,1 and Matthias Schmidt1, ∗
1
Theoretische Physik II, Physikalisches Institut, Universität Bayreuth, D-95447 Bayreuth, Germany
(Dated: 26 January 2023)
Liquid structure carries deep imprints of an inherent thermal invariance against a spatial transfor-
mation of the underlying classical many-body Hamiltonian. At first order in the transformation field
the Noether theorem yields the local force balance. Three distinct two-body correlation functions
emerge at second order, namely the standard two-body density, the localized force-force correlation
function, and the localized force gradient. An exact Noether sum rule interrelates these correlators.
Simulations of Lennard-Jones and Weeks-Chandler-Andersen liquids, of monatomic water and of a
colloidal gel former demonstrate their fundamental role in the characterization of spatial structure.
arXiv:2301.11221v1 [cond-mat.soft] 26 Jan 2023

It is a both surprising and intriguing phenomenon that order level the thermal Noether invariance leads to exact
the liquid phase occurs in the phase diagram, at high identities that form a comprehensive statistical two-body
enough density and low enough temperature, at and off correlation framework. We use simulations to demon-
coexistence with the gas or the solid phase. Famously strate the relevance for the investigation of the structure
it has been argued [1, 2] that it needs relying on ob- of simple, beyond-simple and gelled liquids.
servations rather than mere theory alone to predict the We consider systems of N particles with positions
existence of the liquid state, as neither the noninteract- r1 , . . . , rN ≡ rN and momenta p1 , . . . , pN ≡ pN . The
ing ideal gas nor the Einstein crystal form appropriate Hamiltonian consists of kinetic, interparticle, and exter-
idealized references for a liquid. The difficulty and intri- nal energy contributions,
cacy of the theoretical description of the liquid state [3–6] X p2 X
i
lies in the unusual combination of high spatial symmetry H= + u(rN ) + Vext (ri ), (1)
2m
against global translations and rotations, together with i i
the correlated and strongly interacting behaviour of the where the indices i = 1, . . . , N run over all particles,
dense constituents, whether they are atoms, molecules, m indicates the particle mass, u(rN ) is the interparticle
or colloids. interaction potential, and Vext (r) is a one-body external
Among the defining features of the liquid state are the potential as a function of position r.
ability to spontaneously form an interface when at liquid- We consider a canonical transformation [30], where co-
gas coexistence, the viscous response against shearing ordinates and momenta change according to the following
motion, and the rich pair correlation structure. While map [28, 29]:
the one-body density distribution is homogeneous in bulk ri → ri + (ri ), (2)
(in stark contrast to the microscopic density of a crys-
pi → (1 + ∇i (ri ))−1 · pi . (3)
tal), the joint probability of finding two particles at a
given separation distance r is highly nontrivial in a liq- Here (r) is a spatial “shifting” field that parameterizes
uid. The pair correlation function g(r) [4], as accessible the transform, 1 indicates the 3 × 3-unit matrix and the
e.g. via microscopy techniques [7–9], quantifies this spa- superscript −1 of a matrix is its inverse. The transforma-
tial structure on the particle level. At large distances r, tion (2) and (3) preserves the phase space volume element
the asymptotic decay of g(r) falls into different classes and the Hamiltonian [28–30].
[10–13] with much current interest in electrolytes [14]. We consider the shifting field and its gradient to be
It is a common strategy to exploit the symmetries of small and hence Taylor expand. The coordinate transfor-
a given physical system via Noether’s theorem of invari- mation (2) is already linear in the displacement field and
ant variations [15, 16]. From symmetries in the dynam- is hence unaffected. The momentum transformation (3),
ical description of the system one systematically obtains when expanded as a geometric (Neumann) series to sec-
conservation laws. Typically the starting point is the ond order, is:
action functional, as generalized to a variety of statis- (1 + ∇i (ri ))−1 = 1 − ∇i (ri ) + (∇i (ri ))2 − . . . . (4)
tical mechanical settings [17–24]. In contrast, we have
When expressed in the new variables, the Hamiltonian
recently applied Noether’s concept directly to statistical
acquires a functional dependence on the shifting field,
mechanical functionals, such as the free energy [25–28].
i.e. H → H[]. It is then straightforward to show [28, 29]
This allows to exploit a specific thermal invariance prop-
erty of Hamiltonian many-body systems against shifting that the locally resolved one-body force operator F̂(r)
as performed globally [25–27] or locally resolved in posi- follows from functional differentiation according to:
tion [28, 29]. δH[]
− = F̂(r), (5)
In this Letter we demonstrate that at the local second- δ(r) =0
2

where δ/δ(r) indicates the functional derivative with re-


spect to the shifting field (r). As indicated, (r) is set to k’’
zero after the derivative has been taken. Similar to the
structure of the Hamiltonian (1), the one-body force op- k’ j
erator F̂(r) contains kinetic, interparticle, and external
contributions:

F̂(r) = −∇ ·
X pi pi

i
m
δ(r − ri ) + F̂int (r) − ρ̂(r)∇Vext (r).

(6)
i k
Here δ(·) indicates the (three-dimensional) Dirac distri-
bution, F̂int (r) = − i δ(r−ri )∇i u(rN ) is thePinterparti-
P
cle one-body force operator [31], and ρ̂(r) = i δ(r − ri )
is the standard one-body density operator [4, 5].
We complement this deterministic description by the FIG. 1: Illustration of the three different correlation functions
statistical mechanics of the grand ensemble at chemical that are constrained by thermal Noether invariance. The par-
potential µ and temperature T . The grand potential is ticles (spheres) exert forces (arrows) onto each other. Parti-
Ω = −kB T ln Ξ, with the the grand partition sum Ξ = cles i and j interact directly with each other (black arrows).
Tr e−β(H−µN ) . Here kB indicates the Boltzmann con- The total force (white arrow) on each particle is also deter-
mined by the forces that all other particles k, k0 , k00 exert (pink
stant, β = 1/(kB T ) denotes inverse temperature, and the arrows). The force-force correlations are balanced by the po-
classical “trace”
P∞ operation inRthe grand ensemble is given tential energy curvature ∇i ∇j βu(rN ) (orange surface) and by
3N −1
R
by Tr · = N =0 (N !h ) dr 1 . . . dr N dp1 . . . dpN ·, the two-body density Hessian ∇∇0 ρ2 (r, r0 ) (black curve).
where h denotes the Planck constant. The correspond-
ing grand probability distribution is Ψ = e−β(H−µN ) /Ξ
and thermal averages are defined via h·i = Tr Ψ·, as is
standard. A primary example of a thermal average is the δH[]/δ(r) as the negative force density operator via
density profile being the average of the one-body density Eq. (5), inserting Eq. (9) into Eq. (8), and re-arranging
operator, i.e. ρ(r) = hρ̂(r)i. gives the following locally resolved two-body Noether
Via the transformed Hamiltonian H[], the grand par- sum rule:
tition sum acquires functional dependence on the shifting
field [28, 29], i.e. Ξ[], and so does the grand potential,
i.e. Ω[]. Noether invariance [25, 26] however implies that D δ 2 H[] E
the grand potential does not change under the transfor- βhF̂(r)F̂(r0 )i = . (10)
δ(r)δ(r0 ) =0
mation, and hence

Ω[] = Ω, (7)
We have replaced cov(F̂(r), F̂(r0 )) = hF̂(r)F̂(r0 )i, be-
irrespectively of the form of (r). The first functional cause hF̂(r)i = 0 in equilibrium [28, 29]. The sum rule
derivative of Eq. (7) with respect to the shifting field (r) (10) relates the force-force correlations at two different
then yields [28, 29] the locally resolved equilibribum force positions (left hand side) with the mean curvature of the
balance relation F(r) = hF̂(r)i = 0 [4, 31]. Hamiltonian with respect to variation in the shifting field
Here we work at the second-order level and hence con- (right hand side). That such physically meaningful aver-
sider the second derivative of Eq. (7), which yields ages are related to each other, at all positions r and r0 ,
is highly nontrivial.
δ 2 Ω[]
= 0. (8)
δ(r)δ(r0 ) =0 We can bring the fundamental Noether two-body
sum rule (10) into a more convenient form by multi-
Evaluating the functional derivative on the left hand plying by β, splitting off the trivial kinetic contribu-
side gives tions, and introducing the potential energy force oper-
δ 2 Ω[]  δH[] δH[]  D δ 2 H[] E ator F̂U (r), which combines interparticle and external
= −β cov , + , forces according to F̂U (r) = F̂int (r) − ρ̂(r)∇Vext (r). Fur-
δ(r)δ(r0 ) δ(r) δ(r0 ) δ(r)δ(r0 )
(9) thermore we focus on the distinct contributions (sub-
script “dist”) such that only pairs of particles with un-
where the covariance of two observables (phase space equal indices
PNarePinvolved and double sums reduce to
P N
functions) Â and B̂ is defined in the standard way as ij(6=) ≡ i=1 j=1,j6=i . This allows to identify from
cov(Â, B̂) = hÂB̂i − hÂihB̂i. Rewriting the derivative Eq. (10) the following exact distinct two-body Noether
3

identity: gff ⊥ (r) = 0 due to the absence of a third particle at


D E ρb → 0 that could mediate a transversal force.
β F̂U (r)β F̂U (r0 ) = ∇∇0 ρ2 (r, r0 ) We substantiate this Noether correlation framework
dist
with computer simulations using adaptive Brownian dy-
namics [32], which is an algorithm that is both fast
DX E
+ δ(r − ri )δ(r0 − rj )∇i ∇j βu(rN ) .
ij(6=)
and allows for tight control of force evaluation er-
(11) rors. We investigate the Lennard-Jones (LJ) liquid, the
purely repulsive Weeks-Chandler-Andersen (WCA) liq-
Here the two-body density P is defined as is standard: uid, monatomic water [33, 34], and a three-body colloidal
ρ2 (r, r0 ) = hρ̂(r)ρ̂(r0 )idist = h ij(6=) δ(r − ri )δ(r0 − rj )i. gel former [35, 36]. The results are summarized in Fig. 2
The relationship between the different correlators is and we first discuss the two simple liquids. Both the
graphically illustrated in Fig. 1. LJ and the WCA liquid feature pair correlation func-
We demonstrate that this framework has profound im- tions g(r) that display the familiar strongly structured,
plications already for a bulk liquid, where ρ(r) = ρb = damped oscillatory form [4, 10, 11], with a prominent first
const and Vext (r) = 0, such that F̂U (r) = F̂int (r). In peak indicating a nearest neighbor correlation shell and
view of the form of the distinct sum rule (11), we use subsequent, increasingly washed out oscillations at larger
the pair correlation function g(|r − r0 |) = ρ2 (r, r0 )/ρ2b , distances. In stark contrast, both the force-gradient (po-
and introduce both the force-force pair correlation func- tential curvature) correlator g∇f (r) and the force-force
tion gff (|r − r0 |) = β 2 hF̂int (r)F̂int (r0 )idist /ρ 2 correlator gff (r) have very different forms than g(r) it-
0
Pb , and the self. The curvature correlator has very strongly localized
force gradient correlator g∇f (|r − r |) = −h ij(6=) δ(r −
positive (⊥) and negative (k) peaks near r = σ, with
ri )δ(r0 − rj )∇i ∇j βu(rN )i/ρ2b , which is also the negative
σ denoting the particle size. This feature is due to the
mean potential curvature. The identity (11) can then be
strong first peak of g(r) combined with the properties of
written succinctly as:
φ0 (r) and φ00 (r), as is evident via Eq. (15), which we find
∇∇g(r) + g∇f (r) + gff (r) = 0, (12) to be satisfied to high numerical accuracy. Our results
confirm the expectation [11] that g(r) is hardly affected
where r = |r − r0 | denotes the separation distance be- by interparticle attraction. In contrast the force gradient
tween the two positions. Both gff (r) and g∇f (r) have g∇f ⊥ (r) has a clear and significant peak in the attractive
tensor rank two, i.e. they are 3 × 3-matrices. Due to the region of the LJ potential with no such feature occurring
rotational symmetry of the bulk liquid, the only nontriv- in the purely repulsive WCA liquid.
ial tensor components are parallel (k) and transversal (⊥) The force-force correlator gff (r) has a similar first peak
to r − r0 , such that Eq. (12) reduces to structure as the curvature correlator, but oscillations ex-
tend much further out to larger distances r. Hence gff (r)
g 00 (r) + g∇f k (r) + gff k (r) = 0, (13) captures also the indirect interactions that are mediated
0
g (r)/r + g∇f ⊥ (r) + gff ⊥ (r) = 0, (14) by surrounding particles; we recall Fig. 1. The strong
negative double peak of the parallel component indicates
with the prime(s) denoting the derivative(s) with respect anti-correlated force orientations, which reflect the direct
to r. In the chosen coordinate system the matrices are interactions between pairs of particles. Both tensor com-
diagonal, diag(k, ⊥, ⊥), with the first axis being parallel ponents of gff (r) satisfy the Noether sum rules (13) and
to r − r0 . (14) to excellent numerical accuracy.
For simple fluids, where the particles interact mutually To go beyond simple liquids, we first turn to the
only via a pair potential φ(r), the force gradient correla- monatomic water model by Molinero and Moore [33],
tor reduces to g∇f (r) = βg(r)∇∇φ(r) such that which includes three-body interparticle interactions in
u(rN ) that generate the tetrahedral coordination of liq-
g∇f k (r) = βg(r)φ00 (r), g∇f ⊥ (r) = βg(r)φ0 (r)/r. (15) uid water. The monatomic water model gives a surpris-
ingly accurate description of the properties of real water,
This simplification is due to theP reduction of the mixed see Ref. [34] for very recent work, while the particles re-
derivative ∇i ∇j u(rN ) = ∇i ∇j kl(6=) φ(|rk − rl |)/2 = main spherical and there is no necessity to explicitly in-
∇i ∇j φ(|ri − rj |), for i 6= j, and it allows to rewrite the voke molecular orientational degrees of freedom. Hence
curvature correlator in Eqs. (13) and (14). The validity our framework (13) and (14) applies. The third column of
of the resulting identities g 00 (r)+βφ00 (r)g(r)+gff k (r) = 0 Fig. 2 demonstrates at ambient conditions that while the
and g 0 (r)/r + βφ0 (r)g(r)/r + gff ⊥ (r) = 0 can be analyt- shape of g(r) is similar to that in the LJ liquid, both the
ically verfied in the low-density limit of a simple fluid, potential energy curvature and the force-force correlator
where g(r) = exp(−βφ(r)). Then the force-force corre- differ markedly from those of the LJ model. Notably the
lations are due to the antiparallel direct forces between shape of the double negative peak differs and the sign of
a particle pair: gff k (r) = −g(r)[βφ0 (r)]2 . Furthermore gff ⊥ (r) does not turn negative for distances towards the
4

Lennard-Jones liquid WCA liquid Monatomic water Three-body gel


ρb = 0.844σ −3 , kB T = 0.71ε ρb = 0.844σ −3 , kB T = 0.71ε ρb = 0.458σ −3 , kB T = 0.096ε ρb = 0.037σ −3 , kB T = 0.1ε
S S = 3.5 S = 3.5 S = 2.5 S = 17 g

S
2

0
S S = 1400 S = 1400 S = 400 S = 3500 g∇f k
S = 60 βgφ00 S = 60 βgφ00 S = 35 S = 150
g∇f ⊥
S
βgφ0 /r βgφ0 /r
2
σ −2

− S2

−S
S S = 1400 S = 1400 S = 400 S = 3500 gff k
S = 60 S = 60 S = 35 S = 150
−g 00 − g∇f k
S gff ⊥
2
−g 0 /r − g∇f ⊥
σ −2

− S2

−S
1 2 1 2 1 2 1 2
r/σ r/σ r/σ r/σ

FIG. 2: Simulation results for the two-body correlation functions of the Lennard-Jones liquid (first column), the WCA liquid
(second column), monatomic water (third column), and the three-body gel (fourth column). Results are shown as a function
of the scaled interparticle distance r/σ and S is a scale factor for the vertical axes. Shown is the pair correlation function
g(r) (top row), the potential curvature correlator g∇f (r) (middle row) and the force-force correlator gff (r) (bottom row); the
latter two correlators have a transversal (⊥) and a parallel (k) tensor component. The results for g∇f (r) for the LJ and WCA
liquids are numerically identical to those from the analytical expressions (15) (dashed lines). The directly sampled results for
gff (r) are numerically identical to those obtained from the Noether sum rules (13) and (14) (dashed lines) for all four systems.
Vertical gray lines indicate the position of the first maximum of g(r) as a guide to the eye.

second shell, as is the case in the LJ liquid. Consistently, liquids. While g(r) has the generic long-range decay that
the magnitude of the k components is much larger than one expects of network-forming systems, both the curva-
the ⊥ component, as direct interparticle interactions are ture and the force-force correlator are much more specific
prominent in the former, whereas mediation by third par- indicators. In particular we attribute the striking shape
ticles is required for the latter. of the transversal (⊥) tensor component to the network
connectivity. Again the sum rules are satisfied to very
The three-body gel former by Kob and coworkers [35, good numerical accuracy which we take i) as a demon-
36] alters the prefered angle of the three-body interaction stration that the gel state is indeed equilibrated, which
term from tetrahedral to stretched (we use 180 degrees distinguishes this model [35, 36] from genuine nonequi-
[37]). This change induces an affinity for the formation librium gel formers, and ii) as a confirmation of the gen-
of chains while retaining an ability for their branching erality and fitness of the Noether correlators to system-
and thus the model forms networks in equilibrium. The atically quantify complex spatial structure formation.
results shown in the fourth column of Fig. 2 indicate
markedly different behaviour as compared to the above In conclusion, we have formulated and tested a system-
5

atic two-body correlation framework based on invariance crossover, J. Chem. Phys. 121, 7869 (2004).
against an intrinsic symmetry of thermal many-body sys- [14] P. Cats, R. Evans, A. Härtel, and R. van Roij, Primitive
tems. It remains to point out formal similarities with sum model electrolytes in the near and far field: Decay lengths
rules for interfacial Hamiltonians [38] (as used for studies from DFT and simulations, J. Chem. Phys. 154, 124504
(2021).
of wetting [39]), the concept of an effective temperature [15] E. Noether, Invariante Variationsprobleme, Nachr. d.
[40], the study of one-dimensional systems [41, 42], the König. Gesellsch. d. Wiss. zu Göttingen, Math.-Phys.
possible implications for interfacial physics [34, 43], for Klasse, 235, 183 (1918). English translation by M. A.
the structure of crystals [44–46] and gels [35, 36, 47], Tavel: Invariant variation problems. Transp. Theo. Stat.
for force-sampling simulation techniques [48–50], and for Phys. 1, 186 (1971); for a version in modern typesetting
force-based classical [29, 51] and quantum density func- see: Frank Y. Wang, arXiv:physics/0503066v3 (2018).
tional theory [52–55]. Besides measurements of g(r) [7– [16] N. Byers, E. Noether’s discovery of the deep con-
nection between symmetries and conservation laws,
9], position-resolved forces have recently become accessi- arXiv:physics/9807044 (1998).
ble by direct imaging in colloidal systems [56], which can [17] A. G. Lezcano and A. C. M. de Oca, A stochastic version
facilitate experimental investigations of Noether correla- of the Noether theorem, Found. Phys. 48, 726 (2018).
tors. [18] J. C. Baez and B. Fong, A Noether theorem for Markov
processes, J. Math. Phys. 54, 013301 (2013).
We thank Andrew O. Parry for pointing out Ref. [38] [19] I. Marvian and R. W. Spekkens, Extending Noether’s
to us and him and Gerhard Jung for useful discussions. theorem by quantifying the asymmetry of quantum
states, Nat. Commun. 5, 3821 (2014).
[20] S. Sasa and Y. Yokokura, Thermodynamic entropy as a
Noether invariant, Phys. Rev. Lett. 116, 140601 (2016).
[21] S. Sasa, S. Sugiura, and Y. Yokokura, Thermodynamical

Electronic address: Matthias.Schmidt@uni-bayreuth.de path integral and emergent symmetry, Phys. Rev. E 99,
[1] V. F. Weisskopf, About Liquids, Trans. N. Y. Acad. Sci. 022109 (2019).
38, 202 (1977). [22] Y. Minami and S. Sasa, Thermodynamic entropy as a
[2] R. Evans, D. Frenkel, and M. Dijkstra, From simple liq- Noether invariant in a Langevin equation, J. Stat. Mech.
uids to colloids and soft matter, Phys. Today 72, 38 2020, 013213 (2020).
(2019). [23] M. Revzen, Functional integrals in statistical physics,
[3] J. A. Barker and D. Henderson, What is “liquid”? Un- Am. J. Phys. 38, 611 (1970).
derstanding the states of matter, Rev. Mod. Phys. 48, [24] J. C. Baez, Getting to the bottom of Noether’s theorem,
587 (1976). arXiv:2006.14741
[4] J. P. Hansen and I. R. McDonald, Theory of Simple Liq- [25] S. Hermann and M. Schmidt, Noether’s theorem in sta-
uids, 4th ed. (Academic Press, London, 2013). tistical mechanics, Commun. Phys. 4, 176 (2021).
[5] R. Evans, The nature of the liquid-vapour interface and [26] S. Hermann and M. Schmidt, Why Noether’s theorem ap-
other topics in the statistical mechanics of non-uniform, plies to statistical mechanics, J. Phys.: Condens. Matter
classical fluids, Adv. Phys. 28, 143 (1979). 34, 213001 (2022) (Topical Review).
[6] R. Evans, M. Oettel, R. Roth, and G. Kahl, New devel- [27] S. Hermann and M. Schmidt, Variance of fluctuations
opments in classical density functional theory, J. Phys.: from Noether invariance, Commun. Phys. 5, 276 (2022).
Condens. Matter 28, 240401 (2016). [28] S. Hermann and M. Schmidt, Force balance in thermal
[7] C. P. Royall, A. A. Louis, and H. Tanaka, Measuring quantum many-body systems from Noether’s theorem, J.
colloidal interactions with confocal microscopy, J. Chem. Phys. A: Math. Theor. 55, 464003 (2022).
Phys. 127, 044507 (2007). [29] S. M. Tschopp, F. Sammüller, S. Hermann, M. Schmidt,
[8] A. L. Thorneywork, R. Roth, D. G. A. L. Aarts, and R. and J. M. Brader, Force density functional theory in- and
P. A. Dullens, Communication: Radial distribution func- out-of-equilibrium, Phys. Rev. E 106, 014115 (2022).
tions in a two-dimensional binary colloidal hard sphere [30] H. Goldstein, C. Poole, and J. Safko, Classical Mechanics
system, J. Chem. Phys. 140, 161106 (2014). (Addison–Wesley, New York, 2002). Our generator G is
[9] A. Statt, R. Pinchaipat, F. Turci, R. Evans, and C. P. given as F2 in their notation.
Royall, Direct observation in 3d of structural crossover in [31] M. Schmidt, Power functional theory for many-body dy-
binary hard sphere mixtures, J. Chem. Phys. 144, 144506 namics, Rev. Mod. Phys. 94, 015007 (2022).
(2016) [32] F. Sammüller and M. Schmidt, Adaptive Brownian dy-
[10] R. Evans, J. R. Henderson, D. C. Hoyle, A. O. Parry, namics, J. Chem. Phys. 155, 134107 (2021).
and Z. A. Sabeur, Asymptotic decay of liquid structure: [33] V. Molinero and E. B. Moore, Water modeled as an in-
oscillatory liquid-vapour density profiles and the Fisher- termediate element between carbon and silicon, J. Phys.
Widom line, Mol. Phys. 80, 755 (1993). Chem. B 113, 4008 (2009).
[11] R. Evans, R. J. F. Leote de Carvalho, J. R. Henderson, [34] M. K. Coe, R. Evans, and N. B. Wilding, The coexistence
and D. C. Hoyle, Asymptotic decay of correlations in liq- curve and surface tension of a monatomic water model,
uids and their mixtures, J. Chem. Phys. 100, 591 (1994). J. Chem. Phys. 156, 154505 (2022).
[12] M. Dijkstra and R. Evans, A simulation study of the [35] S. Saw, N. L. Ellegaard, W. Kob, and S. Sastry, Struc-
decay of the pair correlation function in simple fluids, J. tural relaxation of a gel modeled by three body interac-
Chem. Phys. 112, 1449 (2000). tions, Phys. Rev. Lett. 103, 248305 (2009).
[13] C. Grodon, M. Dijkstra, R. Evans, and R. Roth, Decay of [36] S. Saw, N. L. Ellegaard, W. Kob, and S. Sastry, Com-
correlation functions in hard-sphere mixtures: Structural puter simulation study of the phase behavior and struc-
6

tural relaxation in a gel-former modeled by three-body via a macroscopic bond limitation, J. Chem. Phys. 145,
interactions, J. Chem. Phys. 134, 164506 (2011). 074906 (2016).
[37] F. Sammüller, D. der las Heras, and M. Schmidt, In- [48] B. Rotenberg, Use the force! Reduced variance estima-
homogeneous steady shear dynamics of a three-body col- tors for densities, radial distribution functions, and local
loidal gel former, J. Chem. Phys. (to appear in the Special mobilities in molecular simulations, J. Chem. Phys. 153,
Topic on Colloidal Gels) arXiv:2210.07679 150902 (2020).
[38] L. V. Mikhheev and J. D. Weeks, Sum rules for interface [49] D. de las Heras and M. Schmidt, Better than counting:
Hamiltonians, Physica A 177, 495 (1991). Density profiles from force sampling, Phys. Rev. Lett.
[39] A. Squarcini, J. M. Romero-Enrique, and A. O. Parry, 120, 218001 (2018).
Casimir Contribution to the Interfacial Hamiltonian for [50] D. Borgis, R. Assaraf, B. Rotenberg, and R. Vuilleumier,
3D Wetting, Phys. Rev. Lett. 128, 195701 (2022). Computation of pair distribution functions and three-
[40] S. Saw, L. Costigliola, and J. C. Dyre, Configurational dimensional densities with a reduced variance principle,
temperature in active-matter models, arXiv:2204.06819. Mol. Phys. 111, 3486 (2013).
[41] C. W. J. Beenakker, Pair correlation function of the one- [51] F. Sammüller, S. Hermann, and M. Schmidt, Should
dimensional Riesz gas, arXiv:2212.02117 classical density functional theory be based on forces?
[42] A. Flack, S. N. Majumdar, and G. Schehr, An exact A comparative study, arXiv:2212.01780.
formula for the variance of linear statistics in the one- [52] I. V. Tokatly, Time-dependent deformation functional
dimensional jellium model, arXiv:2211.11850 theory, Phys. Rev. B 75, 125105 (2007).
[43] D. G. Triezenberg and R. Zwanzig, Fluctuation theory of [53] C. A. Ullrich and I. V. Tokatly, Nonadiabatic electron
surface tension, Phys. Rev. Lett. 28, 1183 (1972). dynamics in time-dependent density-functional theory,
[44] C. Walz and M. Fuchs, Displacement field and elastic Phys. Rev. B 73, 235102 (2006).
constants in nonideal crystals, Phys. Rev. B 81, 134110 [54] M. M. Tchenkoue, M. Penz, I. Theophilou, M. Ruggen-
(2010). thaler, and A. Rubio, Force balance approach for ad-
[45] J. M. Häring, C. Walz, G. Szamel, and M. Fuchs, Coarse- vanced approximations in density functional theories, J.
grained density and compressibility of nonideal crystals: Chem. Phys. 151, 154107 (2019).
General theory and an application to cluster crystals, [55] W. Tarantino and C. A. Ullrich, A reformulation of time-
Phys. Rev. B 92, 184103 (2015). dependent Kohn-Sham theory in terms of the second time
[46] S.-C. Lin, M. Oettel, J. M. Häring, R. Haussmann, derivative of the density, J. Chem. Phys. 154, 204112
M. Fuchs, and G. Kahl, Direct correlation function of (2021).
a crystalline solid, Phys. Rev. Lett. 127, 085501 (2021). [56] J. Dong, F. Turci, R. L. Jack, M. A. Faers, and C. P.
[47] B. A. Lindquist, R. B. Jadrich, D. J. Milliron, and Royall, Direct imaging of contacts and forces in colloidal
T. M. Truskett, On the formation of equilibrium gels gels, J. Chem. Phys. 156, 214907 (2022).

You might also like